The State of the World’s Animal Genetic Resources for Food and Agriculture 9789251057629

523 121 8MB

English Pages [524] Year 2007

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

The State of the World’s Animal Genetic Resources for Food and Agriculture
 9789251057629

Citation preview

THE STATE OF THE WORLD’s

ANIMAL GENETIC RESOURCES FOR FOOD AND AGRICULTURE

COMMISSION ON GENETIC RESOURCES FOR FOOD AND AGRICULTURE FOOD AND AGRICULTURE ORGANIZATION OF THE UNITED NATIONS Rome, 2007

i

The designations employed and the presentation of material in this information product do not imply the expression of any opinion whatsoever on the part of the Food and Agriculture Organization of the United Nations concerning the legal or development status of any country, territory, city or area or of its authorities, or concerning the delimitation of its frontiers or boundaries. The mention of specific companies or products of manufacturers, whether or not these have been patented, does not imply that these have been endorsed or recommended by the Food and Agriculture Organization of the United Nations in preference to others of a similar nature that are not mentioned. The views expressed in this publication are those of the author(s) and do not necessarily reflect the views of the Food and Agriculture Organization of the United Nations. ISBN 978-92-5-105762-9

All rights reserved. Reproduction and dissemination of material in this information product for educational or other noncommercial purposes are authorized without any prior written permission from the copyright holders provided the source is fully acknowledged. Reproduction of material in this information product for resale or other commercial purposes is prohibited without written permission of the copyright holders. Applications for such permission should be addressed to: Chief Electronic Publishing Policy and Support Branch Communication Division FAO Viale delle Terme di Caracalla, 00153 Rome, Italy or by e-mail to: [email protected] © FAO 2007

Citation: FAO. 2007. The State of the World’s Animal Genetic Resources for Food and Agriculture, edited by Barbara Rischkowsky & Dafydd Pilling. Rome.

ii

Foreword

T

he wise management of the world’s agricultural biodiversity is becoming an ever greater challenge for the international community. The livestock sector in particular is undergoing dramatic changes as large-scale production expands in response to surging demand for meat, milk and eggs. A wide portfolio of animal genetic resources is crucial to adapting and developing our agricultural production systems. Climate change and the emergence of new and virulent animal diseases underline the need to retain this adaptive capacity. For hundreds of millions of poor rural households, livestock remain a key asset, often meeting multiple needs, and enabling livelihoods to be built in some of the world’s harshest environments. Livestock production makes a vital contribution to food and livelihood security, and to meeting the United Nations Millennium Development Goals. It will be of increasing significance in the coming decades. And yet, genetic diversity is under threat. The reported rate of breed extinctions is of great concern, but it is even more worrying that unrecorded genetic resources are being lost before their characteristics can be studied and their potential evaluated. Strenuous efforts to understand, prioritize and protect the world’s animal genetic resources for food and agriculture are required. Sustainable patterns of utilization must be established. Traditional livestock keepers – often poor and in marginal environments – have been the stewards of much of our animal genetic diversity. We should not ignore their role or neglect their needs. Equitable arrangements for benefit-sharing are needed, and broad access to genetic resources must be ensured. An agreed international framework for the management of these resources is crucial. This report is the first global assessment of the status and trends of animal genetic resources, and of the state of institutional and technological capacity to manage these resources. It provides a basis for renewed efforts to ensure that the commitments to the improved management of genetic resources set out in the World Food Summit Plan of Action are realized. It is a milestone in the work of the Commission on Genetic Resources for Food and Agriculture. The support provided by the world’s governments, exemplified by the 169 Country Reports submitted to FAO, has been particularly heartening. I am also greatly encouraged by the contribution that the process of preparing this report has already made to awareness of the topic and to catalysing activity at national and regional levels. However, much remains to be done. The launch of The State of the World’s Animal Genetic Resources for Food and Agriculture at the International Technical Conference on Animal Genetic Resources at Interlaken, Switzerland, must be a springboard for action. I wish to take this opportunity to appeal to the international community to recognize that animal genetic resources are a part of our common heritage that is too valuable to neglect. Commitment and cooperation in the sustainable use, development and conservation of these resources are urgently required.

Jacques Diouf FAO Director-General

iii

Contents Acknowledgements Preface The reporting and preparatory process Executive summary

Part 1

xxi xxv xxvii xxxv

The state of agricultural biodiversity in the livestock sector Introduction SECTION A: ORIGIN AND HISTORY OF LIVESTOCK DIVERSITY

SECTION B:

SECTION C:

5

1 Introduction 2 The livestock domestication process 3 Ancestors and geographic origins of our livestock 4 Dispersal of domesticated animals 5 Transformations in livestock following domestication 6 Conclusions References

5 6 10 14 17 18 19

STATUS OF ANIMAL GENETIC RESOURCES

23

1 Introduction 2 State of reporting 3 Species diversity 3.1 The big five 3.2 Other widespread species 3.3 Species with a narrower distribution 4 Breed diversity 4.1 Overview 4.2 Local breeds 4.3 Regional transboundary breeds 4.4 International transboundary breeds 5 Risk status of animal genetic resources 6 Trends in breed status 6.1 Changes in the number of breeds in the different breed groups 6.2 Trends in genetic erosion 7 Conclusions

23 23 27 28 29 30 31 31 34 35 36 37 44 44 45 48

FLOWS OF ANIMAL GENETIC RESOURCES

51

1 Introduction 2 Driving forces and historical phases in gene flows 2.1 Phase 1: prehistory to the eighteenth century 2.2 Phase 2: nineteenth to mid-twentieth centuries 2.3 Phase 3: mid-twentieth century to the present

51 51 52 53 53

v

SECTION C:

SECTION D:

FLOWS OF ANIMAL GENETIC RESOURCES - continued 3 The big five 3.1 Cattle 3.2 Sheep 3.3 Goats 3.4 Pigs 3.5 Chickens 3.6 Other species 4 Impacts of gene flows on diversity 4.1 Diversity-enhancing gene flow 4.2 Diversity-reducing gene flow 4.3 Diversity-neutral gene flow 4.4 The future References

55 56 61 65 67 69 70 71 71 72 72 73 73

USES AND VALUES OF ANIMAL GENETIC RESOURCES

77

1 2 3 4 5 6 7

Introduction Contribution to national economies Patterns of livestock distribution Food production Production of fibre, skins, hides and pelts Agricultural inputs, transport and fuel Other uses and values 7.1 Savings and risk management 7.2 Sociocultural roles 7.3 Environmental services 8 Roles of livestock for the poor 9 Conclusions References

SECTION E:

ANIMAL GENETIC RESOURCES AND RESISTANCE TO DISEASE 101 1 Introduction 2 Disease resistant or tolerant breeds 2.1 Trypanosomiasis 2.2 Ticks and tick-borne diseases 2.3 Internal parasites 2.4 Foot rot 2.5 Bovine leukosis 2.6 Diseases of poultry 3 Opportunities for within-breed selection for disease resistance 4 Conclusions References

vi

77 77 80 84 86 88 90 90 92 95 97 99 100

101 103 104 104 104 106 107 107 108 110 110

SECTION F:

Part 2

THREATS TO LIVESTOCK GENETIC DIVERSITY

113

1 Introduction 2 Livestock sector trends: economic, social and policy factors 3 Disasters and emergencies 4 Epidemics and disease control measures 5 Conclusions References

113 114 120 126 131 132

Livestock sector trends Introduction SECTION A: DRIVERS OF CHANGE IN THE LIVESTOCK SECTOR

SECTION B:

141

1 Changes in demand 1.1 Purchasing power 1.2 Urbanization 1.3 Consumer taste and preference 2 Trade and retailing 2.1 Flows of livestock and their products 2.2 The rise of large retailers and vertical coordination along the food chain 3 Changing natural environment 4 Advances in technology 5 Policy environment

141 143 143 145 145 145

LIVESTOCK SECTOR’S RESPONSE

153

1 Landless industrialized production systems 1.1 Overview and trends 1.2 Environmental issues 2 Small-scale landless systems 2.1 Overview 2.2 Environmental issues 2.3 Trends 3 Grassland-based systems 3.1 Overview 3.2 Environmental issues 3.3 Trends 4 Mixed farming systems 4.1 Overview 4.2 Environmental issues 4.3 Trends 5 Issues in mixed irrigated systems

155 155 161 163 163 164 164 165 165 166 168 170 170 172 173 174

147 149 149 150

vii

SECTION C:

IMPLICATIONS OF THE CHANGES IN THE LIVESTOCK SECTOR FOR GENETIC DIVERSITY 177 References

Part 3

The state of capacities in animal genetic resources management Introduction SECTION A: INSTITUTIONS AND STAKEHOLDERS

SECTION B:

viii

179

187

1 Introduction 2 Analytical framework 2.1 Stakeholders’ involvement and background at country level 2.2 Assessment of institutional capacities at country level 2.3 Organizations and networks with a potential role in regional and international collaboration 3 Stakeholders, institutions, capacities and structures 3.1 Stakeholder involvement in the State of the World process at country level 3.2 Assessment of institutional capacities at country and regional level 3.3 Organizations and networks with a potential role in subregional, regional and international collaboration 4 Conclusions References Annex

187 187 188 188

STRUCTURED BREEDING PROGRAMMES

215

1 Introduction 2 Species priorities and breeding objectives 2.1 Cattle 2.2 Buffaloes 2.3 Sheep and goats 2.4 Pigs 2.5 Poultry 2.6 Other species 3 Organizational structures 4 Tools and implementation 5 Overview of breeding programmes by region 5.1 Africa 5.2 Asia 5.3 Europe and the Caucasus 5.4 Latin America and the Caribbean 5.5 Near and Middle East 5.6 North America and Southwest Pacific 6 Conclusions and future priorities References Annex

215 216 216 217 218 219 219 219 220 222 225 225 227 229 230 232 232 233 235 236

190 190 190 191 196 201 203 204

SECTION C:

SECTION D:

CONSERVATION PROGRAMMES

243

1 Introduction 2 Global status 3 Stakeholders 3.1 National governments 3.2 Universities and research institutes 3.3 Civil society organizations and breeders’ associations 3.4 Farmers 3.5 Part-time or hobby farmers 3.6 Breeding companies 4 Conservation at species level – status and opportunities 4.1 Cattle 4.2 Sheep 4.3 Goats 4.4 Pigs 4.5 Chickens 4.6 Horses 5 In vivo and in vitro conservation programmes – regional analysis 5.1 Africa 5.2 Asia 5.3 Europe and the Caucasus 5.4 Latin America and the Caribbean 5.5 Near and Middle East 5.6 North America 5.7 Southwest Pacific 6 Opportunities for improving conservation programmes 7 Conclusions and priorities References

243 244 245 245 246 246 246 247 247 247 247 248 249 249 249 250 250 250 253 255 256 258 258 259 260 261 263

REPRODUCTIVE AND MOLECULAR BIOTECHNOLOGY

265

1 Introduction 2 Global overview 3 Africa 4 Asia 5 Europe and the Caucasus 6 Latin America and the Caribbean 7 Near and Middle East 8 North America 9 Southwest Pacific 10 Conclusions References

265 265 266 268 269 271 272 272 272 273 273

ix

SECTION E:

Part 4

LEGISLATION AND REGULATION

275

1 International legal framework – major instruments 1.1 Introduction 1.2 Legal framework for the management of biodiversity 1.3 Access and benefit-sharing 1.4 Legal framework for international trade 1.5 Intellectual property rights 1.6 Legal framework for biosecurity 1.7 Conclusions References 2 Emerging legal issues 2.1 Patenting 2.2 Livestock Keepers’ Rights 3 Regulatory frameworks at regional level 3.1 Introduction 3.2 European Union legislation: an example of a comprehensive regional legal framework 3.3 Conclusions Legislation cited 4 National legislation and policy 4.1 Introduction 4.2 Methods 4.3 Implementation of AnGR-related legislation and programmes 4.4 Country Report analysis 4.5 Conclusions References

275 275 275 277 278 279 280 283 284 285 285 291 291 291

State of the art in the management of animal genetic resources Introduction SECTION A: BASIC CONCEPTS

SECTION B:

x

292 301 302 307 307 307 308 309 332 333

339

1 Animal genetic resources and breeds 2 Management of animal genetic resources 3 Risk status classification References

339 340 342 345

METHODS FOR CHARACTERIZATION

347

1 Introduction 2 Characterization – as the basis for decision-making 3 Tools for characterization 3.1 Surveying 3.2 Monitoring 3.3 Molecular genetic characterization 3.4 Information systems 4 Conclusions References

347 347 350 350 352 354 354 357 358

SECTION C:

SECTION D:

MOLECULAR MARKERS – A TOOL FOR EXPLORING GENETIC DIVERSITY

359

1 Introduction 2 The roles of molecular technologies in characterization 3 Overview of molecular techniques 3.1 Techniques using DNA markers to assess genetic diversity 3.2 Using markers to estimate effective population size 3.3 Molecular tools for targeting functional variation 4 The role of bioinformatics 5 Conclusions References

359 361 362 362 367 367 372 372 375

GENETIC IMPROVEMENT METHODS TO SUPPORT SUSTAINABLE UTILIZATION

381

1 Introduction 2 The context for genetic improvement 2.1 Changing demand 2.2 Diverse production environments 2.3 Increasing recognition of the importance of genetic diversity 2.4 Scientific and technological advances 2.5 Economic considerations 3 Elements of a breeding programme 3.1 Breeding goals 3.2 Selection criteria 3.3 Design of breeding scheme 3.4 Data recording and management 3.5 Genetic evaluation 3.6 Selection and mating 3.7 Progress monitoring 3.8 Dissemination of genetic progress 4 Breeding programmes in high-input systems 4.1 Dairy and beef cattle breeding 4.2 Sheep and goat breeding 4.3 Pig and poultry breeding 5 Breeding programmes in low-input systems 5.1 Description of low-input systems 5.2 Breeding strategies 6 Breeding in the context of conservation 6.1 Methods for monitoring small populations 6.2 Conservation through breeding 7 Conclusions References

381 381 381 382 382 382 387 388 390 391 392 392 393 394 395 395 396 396 400 402 405 405 406 420 420 421 421 423

xi

SECTION E:

SECTION F:

xii

METHODS FOR ECONOMIC VALUATION

429

1 Introduction 2 Development of methodologies for economic analysis 3 Application of economic methodologies in animal genetic resources management 3.1 Value of animal genetic resources to farmers 3.2 Costs and benefits of conservation 3.3 Targeting of farmers for participation in in situ breed conservation programmes 3.4 Priority setting in livestock conservation programmes 3.5 Priority setting in livestock breeding strategies 3.6 General policy analysis 4 Implications for policies and future research References

429 431 433 433 434 436 437 438 438 439 440

METHODS FOR CONSERVATION

443

1 Introduction 2 Arguments for conservation 2.1 Arguments related to the past 2.2 Safeguarding for future needs 2.3 Arguments related to the present situation 3 The unit of conservation 4 Conservation of plant versus animal genetic resources 5 Information for conservation decisions 6 In vivo conservation 6.1 Background 6.2 Genetic management of populations 6.3 Self-sustaining strategies for local breeds 6.4 In situ versus ex situ approaches to in vivo conservation 7 Current status and future prospects for cryoconservation 7.1 Gametes 7.2 Embryos 7.3 Cryoconservation of somatic cells and somatic cell cloning 7.4 Choice of genetic material 7.5 Security in genebanks 8 Resource allocation strategies in conservation 8.1 Methods for setting priorities 8.2 Optimization strategies for planning conservation programmes 9 Conclusions References

443 444 445 445 447 448 449 450 454 454 454 455 460 461 462 463 463 465 465 466 466 467 470 473

SECTION G:

Part 5

RESEARCH PRIORITIES

477

1 2 3 4 5 6 7 8 9

477 477 478 478 478 479 480 480 481

Information for effective utilization and conservation Information systems Molecular methods Characterization Genetic improvement methods Conservation methods Decision-support tools for conservation Economic analysis Access and benefit-sharing

Needs and challenges in animal genetic resources management Introduction SECTION A: KNOWLEDGE OF ANIMAL GENETIC DIVERSITY: CONCEPTS, METHODS AND TECHNOLOGIES

SECTION B:

SECTION C:

SECTION D:

487

CAPACITY IN ANIMAL GENETIC RESOURCES MANAGEMENT 493 1 Capacity in characterization, sustainable use and conservation of animal genetic resources 2 Capacity in institutions and policy-making

493 494

MAJOR CHALLENGES FOR LIVESTOCK DEVELOPMENT AND ANIMAL GENETIC RESOURCES MANAGEMENT

499

ACCEPTING GLOBAL RESPONSIBILITY

503

Abbreviations and acronyms

505

Annexes (on CD-ROM) Country Reports Reports from International Organizations Subregional Reports Thematic Studies List of breeds documented in the Global Databank for Animal Genetic Resources List of breeds at risk List of authors, reviewers and their affiliations

xiii

BOXES 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42

xiv

The domestication process Molecular characterization – a tool to understand livestock origin and diversity The history of African pastoralism What is new compared to the World Watch List for Domestic Animal Diversity? Glossary: populations, breeds, regions Glossary: risk status classification Gene flows resulting from colonization Nelore cattle Continuous repackaging of genes – Dorper sheep Hybrid pigs The chicken breeding industry Linguistic links between cattle and wealth The history of Hungarian Grey cattle – changing uses over time Genetic resistance to African swine fever Mongolian reindeer under threat Policy distortions influencing the erosion of pig genetic resources in Viet Nam Which dairy breeds for tropical smallholders? War and rehabilitation in Bosnia and Herzegovina The concept of productivity Sustainable utilization of the Iberian pig in Spain – a success story Overcoming constraints to the development of small-scale market-oriented dairying Facts and trends in the emerging world food economy Suggestions for strengthening national structures Research and breed development in Africa Sheep breeding in Tunisia Buffalo breeding in India Goat breeding in the Republic of Korea Duck breeding in Viet Nam Pig breeding in Hungary Horse breeding – tradition and new requirements Beef cattle breeding in Brazil Breeding llamas in Argentina Influence of market forces on livestock breeding in the United States of America Sheep breeding in Australia Mali – role of the government Ethiopia – in situ conservation Morocco’s Plan Moutonnier – designated breeding areas to sustain local sheep breeds Conservation strategies in China Denmark – opportunities for in vivo conservation Brazil – implementation of a genebank United States of America – priorities in conservation programmes Australia – involvement of diverse stakeholders

6 9 15 24 25 37 53 60 64 67 70 91 96 109 116 118 119 126 140 144 146 151 202 226 226 227 228 228 230 230 231 232 233 233 246 251 252 254 256 258 259 260

43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 76 77 78 79 80 81 82 83 84 85

Impact of international zoosanitary regulations on animal genetic resources management – the example of FMD The first patented animal The African Union Model Law Malawi’s Environmental Management Act Turkey’s Law on Pastures No. 4342 (1998) Slovenia’s Livestock Breeding Act (2002) Mozambique’s Livestock Development Policy and Strategies Slovenia’s regulation on Conservation of Farm Animal Genetic Resources Uganda’s National Animal Genetic Resources Progamme Ukraine’s Law on Animal Breeding Turkey’s Regulation on Protection of Animal Genetic Resources (2002) Lesotho’s Importation and Exportation of Livestock and Livestock Products Proclamation Malaysia’s Animals Ordinance Hungary’s Decree No. 39 Botswana’s Stock Diseases (Semen) Regulations Barbados’s incentive programme Uganda’s Animal Breeding Act (2001) Guatemala – decentralization of the registration of pure-bred animals Mongolia’s White Revolution Programme The Philippine’s White Revolution Russian Federation – Veterinary and Sanitary Requirements No. 13-8-01/1-8 (1999) India – rules for transportation West Africa – pastoralists crossing borders The Islamic Republic of Iran’s Act of National Veterinary System (1971) Definition of breed adopted by FAO Production environment descriptors for animal genetic resources Information systems at global level DNA, RNA and protein The new “–omics” scientific disciplines Recent developments in molecular biology Extraction and multiplication of DNA and RNA Commonly used DNA markers Sampling genetic material QTL mapping The population genomics approach Databases of biological molecules Glossary: molecular markers Changing body size of beef cattle in the United States of America Calving problems in Belgian White Blue cattle Cross-breeding to address inbreeding-related problems in Holstein cattle Norwegian Red Cattle – selection for functional traits Community-based sheep management in the Peruvian Andes Genetic improvement of an indigenous livestock breed – Boran cattle in Kenya

282 286 292 308 310 311 312 314 315 316 316 318 318 319 320 320 322 323 324 325 327 328 328 330 339 350 356 359 360 360 362 363 364 367 369 372 374 391 396 397 399 406 407

xv

86 A llama breeding programme in Ayopaya, Bolivia 87 Pastoralists’ breeding criteria – insights from a community member 88 The Bororo Zebu of the WoDaaBe in Niger – selection for reliability in an extreme environment 89 Community-driven breeding programmes for local pig breeds in north Viet Nam 90 The cost of heterosis 91 Nigeria’s Village Poultry Improvement Scheme 92 A community-based and participatory dairy goat cross-breeding programme in a low-input smallholder system in the eastern highlands of Kenya 93 Economic values 94 Glossary: conservation 95 Red Maasai sheep – accelerating threats 96 Lleyn sheep of Wales – revival in fortunes in tune with modern demands 97 Decision-making in conservation and utilization – use of genetic diversity data 98 Spatial analysis of genetic diversity 99 In situ conservation of the Norwegian Feral Sheep 100 Examples of incentive payment schemes at the national level 101 An index of economic development potential for targeting in situ conservation investments 102 Community-based in situ conservation programme – a case from Patagonia 103 Changes in production systems leading to replacement of local buffaloes – a case from Nepal 104 Revival of the native Red and White Friesian cattle in the Netherlands 105 Revival of the Enderby cattle in New Zealand 106 Glossary: objective decision aids 107 Optimum allocation of conservation funds – an example featuring African cattle breeds 108 The Svalbard Global Seed Vault: an international seed depository in the Arctic

408 410 412 414 417 417 418 430 443 444 446 452 453 456 457 458 459 460 464 465 468 469 472

TABLES

1 2 3 4 5 6 7 8

9 10 11

xvi

Regional overview of Country Reports Country Reports received Reports from international organizations Origin and domestication of livestock species Status of information recorded in the Global Databank for Animal Genetic Resources Distribution of mammalian species by region Distribution of avian species by region Proportion of the world’s population size (2005) and number of local and regional transboundary breeds (January 2006) of the major livestock species by region Mammalian species – number of reported local breeds Avian species – number of reported local breeds Mammalian species – number of reported regional transboundary breeds

xxviii xxix xxx 7 23 26 27

33 34 34 35

12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48

Avian species – number of reported regional transboundary breeds Mammalian species – number of reported international transboundary breeds Avian species – number of reported international transboundary breeds Number of extinct mammalian breeds Number of extinct avian breeds Years of extinction Reclassification of regional and international transboundary breeds from 1999 to 2006 Changes in risk status of transboundary breeds from 1999 to 2006 Risk status of transboundary breeds reported after 1999 Changes in risk status of local breeds (1999) reclassified as transboundary breeds (2006) Changes in risk status of local breeds from 1999 to 2006 Risk status of local breeds reported after 1999 Workforce employed in agriculture and land area per agricultural worker Number of animals by species/1000 human population Number of animals by species/1000 ha agricultural land Production of food of animal origin (kg/person/year) Production of fibres, skins and hides (1000 tonnes/year) Trends in the use of animals for draught power Roles of livestock by livelihood strategy Selected studies indicating breed difference in resistance or tolerance to specific diseases Mammalian breeds reported to DAD-IS as having resistance or tolerance to specific diseases or parasites Breeds reported to DAD-IS as showing resistance or tolerance to trypanosomiasis Breeds reported to DAD-IS as showing resistance or tolerance to tick-burden Breeds reported to DAD-IS as showing resistance or tolerance to tick-borne diseases Breeds reported to DAD-IS as showing resistance or tolerance to internal parasites Breeds reported to DAD-IS as showing resistance or tolerance to foot rot Cattle breeds reported to DAD-IS as showing resistance or tolerance to leukosis Breeds reported to DAD-IS as showing resistance or tolerance to avian diseases Impact of recent disease epidemics Examples of breeds affected by the FMD outbreak in the United Kingdom in 2001 Projected trends in meat consumption from 2000 to 2050 Projected trends in milk consumption from 2000 to 2050 Standards in the livestock market and implications for small-scale producers Trends in production of meat and milk in developing and developed countries Livestock numbers and production of the world’s livestock production systems – averages for 2001-2003 The developing countries with the highest meat and milk production (2004) Agriculture’s contribution to global greenhouse gas and other emissions

36 36 36 43 43 43 45 46 46 46 47 47 79 83 83 84 87 88 98 102 103 104 105 105 106 107 107 108 128 129 142 143 148 155 157 157 162

xvii

49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 76 77 78 79 80 81 82 83 84 85 86 87 88 89 90 91

xviii

Estimated number of pastoralists in different geographic regions Land with rainfed crop production potential Main crop–animal interactions in crop-based livestock systems Share of irrigated production in total crop production of developing countries Sources of information (Country Report sections) for the national-level assessments Institutional assessment – infrastructure, capacities and participation Institutional assessment – research and knowledge Institutional assessment – state of policy development Organizations and networks that play or may play a role in animal genetic resources management at regional/subregional level Institutional assessment at country level List of international organizations and reports on their activities Countries prioritizing breeding activities (by species) Structured breeding activities for the main livestock species Strategies and tools used in cattle breeding Training, research and farmers’ organizations in current policies Stakeholder involvement in the development of animal genetic resources Number of countries reporting the use of artificial insemination Importance of species and locally adapted versus exotic breeds in current policies List of subsample countries that provided information in predefined tables Strategies and tools used in sheep breeding Strategies and tools used in goat breeding Strategies and tools used in pig breeding Strategies and tools used in chicken breeding Countries reporting structured breeding activities in minor species Stakeholder involvement in structured cattle breeding activities Stakeholder involvement in structured sheep breeding activities Stakeholder involvement in structured goat breeding activities Stakeholder involvement in structured pig breeding activities Number of countries with conservation programmes Conservation activities – at the global level Conservation activities in Africa Conservation activities in Asia Conservation activities in Europe and the Caucasus Conservation activities in Latin America and the Caribbean Conservation activities in the Near and Middle East Conservation activities in North America Conservation activities in the Southwest Pacific Use of biotechnologies by region Use of biotechnologies by species Instruments for sustaining livestock production systems Instruments in the field of conservation Instruments in the field of genetic improvement Instruments related to institutions active in genetic improvement

166 171 171 175 189 192 193 194 197 207 214 217 217 218 220 222 224 224 236 237 237 238 239 239 240 240 241 241 245 248 251 253 256 257 258 259 260 265 266 312 315 317 322

92 93 94 95 96 97 98 99 100 101 102 103 104 105

Instruments in the field of standard setting Instruments for promoting trade in livestock products Instruments regulating import and export of genetic material Instruments regulating livestock movements and import and export of live animals and livestock products Regulations in the field of animal health Information recorded for mammalian species in the Global Databank for Animal Genetic Resources Information recorded for avian species in the Global Databank for Animal Genetic Resources Breeding objectives in ruminants Breeding objectives in pigs Breeding objectives in poultry Overview of valuation methodologies Conservation benefits and costs under a range of valuation methodologies – the case of the Box Keken pig (Yucatan, Mexico) Comparisons of biological, operational and institutional factors influencing plant and animal genetic resources conservation Current status of cryoconservation techniques by species

323 326 326 329 330 351 352 397 402 403 432 435 451 464

FIGURES

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

Assignment of countries to regions and subregions in this report xxxiii Archaeological map of agricultural homelands and spread of Neolithic/Formative cultures, with approximate radiocarbon dates 5 Major centres of livestock domestication – based on archaeological and molecular genetic information 10 Origin and migration routes of domestic cattle in Africa 16 Proportion of national breed populations for which population figures have been reported 25 Regional distribution of major livestock species in 2005 28 Distribution of the world’s mammalian breeds by species 28 Distribution of the world’s avian breeds by species 29 Number of local and transboundary breeds at global level 32 Number of local and transboundary breeds at regional level 32 Proportion of the world’s breeds by risk status category 38 Risk status of the world’s mammalian breeds in January 2006: absolute (table) and percentage (chart) figures by species 39 Risk status of the world’s avian breeds in January 2006: absolute (table) and percentage (chart) figures by species 40 Risk status of the world’s mammalian breeds in January 2006: absolute (table) and percentage (chart) figures by region 41 Risk status of the world’s avian breeds in January 2006: absolute (table) and percentage (chart) figures by region 42

xix

16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42

43 44 45 46 47 48

xx

Local, regional and international breeds in 1999 and 2006 Changes in risk status of transboundary breeds from 1999 to 2006 Changes in risk status of local breeds from 1999 to 2006 Distribution of transboundary breeds Distribution of Holstein-Friesian cattle Distribution of Charolais cattle Distribution of transboundary cattle breeds of Latin American, African or South Asian origin Distribution of transboundary sheep breeds Gene flow of improved Awassi and Assaf sheep from Israel Distribution of Saanen goats Distribution of Boer goats Distribution of Large White pigs Contribution of agriculture and livestock to total GDP by region Contribution of livestock to agricultural GDP Percentage of permanent pasture in total agricultural land Livestock density in relation to human population Livestock density per square kilometre of agricultural land Net exports – meat Net exports – milk equivalent Net exports – eggs Number of disasters by type and year Changes in the meat consumption of developing and developed countries Distribution of livestock production systems Meat production from ruminants versus monogastrics in developing and developed countries Changes in the quantity of cereals used as feed (1992-1994 and 2020) Changes in the distribution of the size of pig farms in Brazil (1985 to 1996) Estimated contribution of livestock to total phosphate supply on agricultural land in areas presenting a phosphate mass balance of more than 10 kg per hectare in selected Asian countries (1998 to 2000) State of institutions – regional comparison State of institutions – subregional comparison within Africa State of institutions – subregional comparison within Asia State of institutions – subregional comparisons within Latin America and the Caribbean Information required to design management strategies Structure of the poultry breeding industry

44 45 47 56 57 57 59 62 64 66 66 68 78 79 80 81 82 85 85 86 121 141 154 156 158 159

161 195 205 205 206 348 389

Acknowledgements

T

his report could not have been prepared without the assistance of the many individuals who generously contributed their time, energy and expertise. FAO would like to take this opportunity to acknowledge these contributions. The core of the information for the State of the World’s Animal Genetic Resources for Food and Agriculture was provided by the 169 governments that submitted Country Reports; the first and most important acknowledgement therefore goes to these governments and to all those individuals in each country who contributed to these reports, in particular the National Coordinators for the Management of Animal Genetic Resources and the National Consultative Committees. The development of training materials the conduct of training workshops, the preparation and analysis of the Country Reports, the follow-up workshops and the various international, regional and national consultations were facilitated by the following team: Daniel Benitez-Ojeda, Harvey D. Blackburn, Arthur da Silva Mariante, Mamadou Diop, M’Naouer Djemali, Anton Ellenbroek, Erling Fimland, Salah Galal, Andreas Georgoudis, Peter Gulliver, Sipke-Joost Hiemstra, Yusup Ibragimov, Jarmo Juga, Ali Kamali, Sergeij Kharitonov, Richard Laing, Birgitta Malmfors, Moketal Joel Mamabolo, Peter Manueli, Elzbieta Martyniuk, Carlos Mezzadra, Rafael Morales, Ruben Mosi, Siboniso Moyo, David R. Notter, Rafael Núñez-Domínguez, Dominique Planchenault, Geoffrey Pollott, Adrien Raymond, Peter Saville, Hermann Schulte-Coerne, Louise Setshwaelo, Paul Souvenir Zafindrajaona, David Steane, Arunas Svitojus, Lutfi Tahtacioglu, Vijay Taneja, Frank Vigh-Larsen, Hans-Gerhard Wagner, Mateusz Wieczorek, Hongjie Yang and Milan Zjalic. An FAO–WAAP (World Association for Animal Production) agreement assisted a large number of developing countries in report preparation. This important contribution to the reporting process could not have been accomplished without the coordination and hard work of Jean Boyazoglu and his colleagues at WAAP. The State of the World’s Animal Genetic Resources for Food and Agriculture was prepared and coordinated by Barbara Rischkowsky with the assistance of Dafydd Pilling. The preparation was facilitated and supported by the Service Chief of Animal Production, Irene Hoffmann, and current and former officers of the Animal Genetic Resources Group: Badi Besbes, David Boerma, Ricardo Cardellino, Mitsuhiro Inamura, Pal Hajas, Keith Hammond, Manuel Luque Cuesta, Beate Scherf, Kim-Anh Tempelman and Olaf Thieme. Administrative and secretarial support was provided by Carmen Hopmans and Kafia FassiFihri. The finalization, layout and printing were supervised by Beate Scherf. The sections of the report were prepared and reviewed by individual experts or expert teams who will be acknowledged below by section. This form of acknowledgement intends to thank the authors for contributing their time, expertise and energy, both in the process of writing, and in reviewing and editing. It will also allow the interested public to identify resource persons for specific topics. This is facilitated by an alphabetical list of authors and reviewers on the attached CD-ROM. Case studies were prepared by: Camillus O. Ahuya, Tony Bennett, Ismaïl Boujenane, Achilles Costales, Erling Fimland, Cary Fowler, John Gibson, Alexander Kahi, John M. King, Saverio Krätli, Maria Rosa Lanari, Ute Lemke, Thomas Loquang, Manuel Luque Cuesta, Paolo Ajmone Marsan, André Markemann, Marnie Mellencamp, Okeyo Mwai, Kor Oldenbroek, John Bryn Owen, Vincente Rodríguez-Estévez, Hans Schiere, Marianna Siegmund-Schulze, Henner Simianer, David Steane, Angelika Stemmer, Kim-Ahn Tempelman, Hongjie Yang and Anne Valle Zárate.

xxi

Additional material for the preparation of text boxes was provided by Brian Donahoe, Morgan Keay, Juhani Mäki-Hokkonen, Kirk Olson and Dan Plumley. Data entry into the Global Databank was carried out by Ellen Geerlings and Lucy Wigboldus. Analysis of the Global Databank was performed by Mateusz Wieczorek, Alberto Montironi, Justyna Dybowska, Kerstin Zander and Beate Scherf. All maps (if not otherwise stated) were prepared by Thierry Lassueur with support from Tim Robinson and Pius Chilonda. Thematic studies were coordinated by Beate Scherf and Irene Hoffmann and prepared by: Erika Alandia Robles, Simon Anderson, Kassahun Awgichew, Roswitha Baumung, P.N. Bhat, Stephen Bishop, Kwame Boa-Amponsem, Ricardo Cardellino, Arthur da Silva Mariante, Mart de Jong, Adam G. Drucker, Christian Gall, Michael Goe, Elisha Gootwine, Douglas Gray, Claire Heffernan, Sipke-Joost Hiemstra, Sabine Homann, Christian G. Hülsebusch, Le Thi Thanh Huyen, Antonella Ingrassia, Ute Lemke, Nils Louwaars, Daniele Manzella, Jacobus Hendrik Maritz, Elzbieta Martyniuk, Marcus Mergenthaler, Klaus Meyn, Giulietta Minozzi, H. Momm, Katinka Musavaya, David R. Notter, Kor Oldenbroek, Marta Pardo Leal, Roswitha Roessler, Cornelia Schäfer, Kim-Anh Tempelman, Morton W. Tvedt and Anne Valle Zárate. Subregional factsheets presented on the attached CD-ROM were prepared by Marieke Reuver, Marion De Vries, Harvey Blackburn, Campbell Davidson, Salah Galal, Ellen Geerlings and National Coordinators from Europe and the Caucasus. Reports on subregional priorities were compiled by Milan Zjalic. Graphic design and layout was by Omar Bolbol and Daniela Scicchigno. Listing every person by name is not easy, and carries with it the risk that someone may be overlooked. Apologies are conveyed to anyone who may have provided assistance whose name has been inadvertently omitted. Any errors or omissions in this work are the responsibility of those who compiled it. None of the contributors should be considered responsible for such defects. In this regard, FAO appreciates any corrections.

xxii

Part / Section

Authors

Reviewers

PART 1: The state of agricultural biodiversity in the livestock sector Origin and history of livestock diversity

Olivier Hanotte

Ilse Koehler-Rollefson

Status of animal genetic resources

Barbara Rischkowsky, Dafydd Pilling, Beate Scherf

Mateusz Wieczorek

Flows of animal genetic resources

Evelyn Mathias, Ilse Koehler-Rollefson, Paul Mundy

Beate Scherf, Annette von Lossau

Uses and values of animal genetic resources

Dafydd Pilling, Barbara Rischkowsky with Manuel Luque Cuesta

Animal genetic resources and disease resistance

Dafydd Pilling, Barbara Rischkowsky

Steve Bishop, Jan Slingenbergh

Threats to livestock genetic diversity

Dafydd Pilling, Claire Heffernan, Michael Goe

Anni McLeod, Simon Mack, Jan Slingenbergh

Pierre Gerber, Dafydd Pilling, Barbara Rischkowsky

Hans Schiere

PART 2: Livestock sector trends

PART 3: The state of capacities in animal genetic resources management Institutions and stakeholders

Maria Brockhaus

Irene Hoffmann, Beate Scherf, Ricardo Cardellino, Jean Boyazoglu, Annette von Lossau, Ilse Koehler-Rollefson

Structured breeding programmes

Olaf Thieme

Juhani Mäki-Hokkonen

Conservation programmes

Kor Oldenbroek with Milan Zjalic

Reproductive and molecular biotechnology

Dafydd Pilling with Milan Zjalic

Salah Galal

International legal framework – major instruments

Dafydd Pilling drawing on FAO legislative study No 89

Clive Stannard, Niels Louwaars

Patenting – an emerging legal issue

Dafydd Pilling with Claudio Chiarolla

Niels Louwaars, Morten Walløe Tvedt

Regulatory frameworks at regional level

Dafydd Pilling drawing on FAO legislative study No 89

Sipke Joost Hiemstra, Danielle Manzella, Hermann Schulte-Coerne, Kai-Uwe Sprenger

National legislation and policy

Susette Biber-Klemm with Cari Rincker

Legislation and regulation

xxiii

Part / Section

Authors

Reviewers

PART 4: The state of the art in animal genetic resources management Basic concepts

Barbara Rischkowsky, Dafydd Pilling

Beate Scherf, Ricardo Cardellino

Methods for characterization

Workneh Ayalew, Beate Scherf, Barbara Rischkowsky

Ed Rege

Molecular markers – a tool for exploring genetic diversity

Paolo Ajmone Marsan, with Kor Oldenbroek

Han Jianlin, Paul Boettcher

Genetic improvement methods to support sustainable utilization

Badi Besbes, Victor Olori, Jim Sanders

Beate Scherf, Ricardo Cardellino, Keith Hammond

Methods for economic valuation

Adam Drucker

Gianni Cicia

Methods for conservation

Jean-Pierre Brillard, Gustavo Gandini, John Gibson, David Notter, Dafydd Pilling, Barbara Rischkowsky, Henner Simianer

Workneh Ayalew, Harvey Blackburn, Jean Boyazoglu, Ricardo Cardellino, Coralie Danchin, Sipke Joost Hiemstra, Elzbieta Martyniuk, Roger Pullin, Beate Scherf, Michele Tixier-Boichard

Research priorities

all authors

all reviewers

PART 5: Needs and challenges in animal genetic resources management Barbara Rischkowsky, Irene Hoffmann

xxiv

Animal Genetic Resources Group and CGRFA Secretariat

Preface

A

gricultural biodiversity is the product of thousands of years of activity during which humans have sought to meet their needs in a wide range of climatic and ecological conditions. Well-adapted livestock have been an essential element of agricultural production systems, particularly important in harsh environments where crop farming is difficult or impossible. The capacity of agro-ecosystems to maintain and increase their productivity, and to adapt to changing circumstances, remains vital to the food security of the world’s population. For livestock keepers, animal genetic diversity is a resource to be drawn upon to select stocks and develop new breeds. More broadly, genetically diverse livestock populations provide society with a greater range of options to meet future challenges. The Food and Agriculture Organization of the United Nations (FAO) has, since the early 1960s, provided assistance to countries to characterize their animal genetic resources for food and agriculture (AnGR) and develop conservation strategies. In 1990, FAO’s Council recommended the development of a comprehensive programme for the sustainable management of AnGR at the global level. A meeting of experts in 1992, and subsequent sessions of FAO’s governing bodies, provided impetus to the development of the Global Strategy for the Management of Farm Animal Genetic Resources, which was initiated in 1993. The Animal Production and Health Division of FAO was designated as the Global Focal Point for Animal Genetic Resources, and given the role of coordinating further development of the Global Strategy. In 1995, the Twenty-eighth Session of the FAO Conference took the decision to broaden the mandate of the Commission on Plant Genetic Resources to cover all aspects of agro-biodiversity of relevance to food and agriculture; the Commission, originally established in 1983, was the first permanent intergovernmental forum dealing with agricultural genetic resources. Work on AnGR was the first element of this expanded role. The Commission was renamed the Commission on Genetic Resources for Food and Agriculture (CGRFA).

The international agenda FAO’s commitment to maintaining agricultural biodiversity is consistent with the increasing prominence of biodiversity on the agenda of the international community. This development is the result of a recognition that threats to biodiversity are increasing, whether measured in terms of the extinction of species, the destruction of ecosystems and habitats, or the loss of genetic diversity within the species utilized for agriculture. The 1992 United Nations Conference on Environment and Development (Earth Summit) held in Rio de Janeiro was an important landmark. The Convention on Biological Diversity (CBD), signed in Rio by 150 governments, committed the nations of the world to conserve their biodiversity, to ensure its sustainable use, and to provide for equitable sharing of the benefits arising from its use. By 2005, 188 countries had become Parties to the CBD. The Conference of Parties (COP) of the CBD (the governing body of the Convention) has specifically recognized the special nature of agricultural biodiversity and the need for distinctive solutions in this field (see for example decision V/5, taken at the Fifth Meeting of the COP in 2000). Agenda 21, adopted by 179 governments at the time at Rio Earth Summit in 1992, is a plan of action to be undertaken at global, national and local levels by governments, the organizations of the United Nations System and other stakeholders, to address all areas of human impact on the environment. The Agenda’s Chapter 14, “Promoting Sustainable Agriculture and Rural Development”, addresses the question of increasing food production

xxv

and enhancing food security in a sustainable way. It included programme areas related to the conservation and development of AnGR. The threat to food security posed by the loss of biodiversity was noted in the Plan of Action adopted at the 1996 World Food Summit held in Rome. Under Objective 3.2(f) of the Rome Declaration, the governments of the world affirmed that they would “promote the conservation and sustainable utilization of animal genetic resources.” Meeting the Millennium Development Goals, adopted by the United Nations in 2000, presents another great challenge to the international community. The adverse effects of biodiversity loss on progress towards the achievement of these goals are cause for concern (UNDP, 2002)1. As well as underpinning food security, biological diversity is the basis of many economic activities, and is vital to ecosystem functioning. Declining biodiversity tends to be associated with greater shocks and fluctuations in ecosystems, and it is the poor that are usually the most vulnerable to these effects. Many poor people are closely dependent on natural resources for their livelihoods, and frequently have a wealth of knowledge regarding the plants and animals with which they work. It has been suggested that this knowledge could be a source of income for the poor if it leads to the development and marketing of unique biological products. In reality, the extent to which the benefits of such developments actually accrue to the poor is often limited – highlighting the need not only for conservation of biodiversity, but for equitable frameworks for its utilization. Within the international framework for the management and conservation of biological diversity, the work of CGRFA focuses on the particular features and problems associated with the management of agro-biodiversity, and the need for distinctive solutions for this field.

1

UNDP. 2002. Building on hidden opportunities to achieve the Millenium Development Goals. Poverty reduction through sustainable biodiversity use, by I Koziell & C.I. McNeill. New York.

xxvi

The reporting and preparatory process

I

n 1999, the CGRFA during its Eighth Regular Session agreed that FAO should coordinate the preparation of a country-driven report on the State of the World’s Animal Genetic Resources for Food and Agriculture (SoW-AnGR)2. In 2004, the Intergovernmental Technical Working Group on Animal Genetic Resources (ITWG-AnGR) – a subsidiary body established by the Commission to address issues relevant to the conservation and sustainable use of AnGR – reviewed progress in the preparation of the SoW-AnGR and endorsed a draft outline including a Report on Strategic Priorities for Action. The CGRFA subsequently endorsed this outline at its Tenth Regular Session. The agreed timetable for the preparation of the SoW-AnGR was that a draft would be available for review by the CGRFA at its Eleventh Regular Session in 2007, and that the report would be finalized at the first International Technical Conference on Animal Genetic Resources. The first draft of the SoW-AnGR was made available to the Fourth Session of the ITWGAnGR in December 2006. The Working Group requested more time to undertake a review of the report. It was agreed that members of the Working Group would provide comments on the draft to FAO by 31 January 2007, in order for FAO to undertake any necessary revisions prior to the presentation of the SoW-AnGR to the CGRFA at its Eleventh Regular Session. The Working Group further agreed that the review process should be open to all Member Countries of the Commission. FAO, therefore, invited all CGRFA Member Countries to submit comments within the agreed time frame.

Inputs to the State of the World’s Animal Genetic Resources reporting process The process of preparing the SoW-AnGR included a number of elements through which the information required was gathered and analysed.

Country Reports In order to ensure the country-driven nature of the process, in March 2001, FAO invited 188 countries to submit Country Reports assessing their AnGR. Guidelines for the preparation of the Country Reports were produced, including a proposed structure. Regional training and follow-up workshops were conducted between July 2001 and November 2004. The overall objectives of the Country Reports were to analyse and report on the state of AnGR, on the status and trends of these resources, and on their current and potential contribution to food, agriculture and rural development; to assess the state of countries’ capacity to manage AnGR, in order to determine priorities for future capacity-building; and to identify national priorities for action in the field of conservation and sustainable utilization of AnGR, and related requirements for international cooperation. The first Country Reports were received in the second half of 2002, with the majority being submitted during 2003 and 2004. The latest Country Report was submitted in October 2005, bringing the total to 169 (Tables 1 and 2). The fact that the submission of the Country Reports was spread over several years meant that as the process of preparing the SoW-AnGR progressed, more information became available for analysis. For this reason, it should be noted that the latest arrivals among the Country Reports could not be fully included in the process of analysis and report preparation. 2

The term animal genetic resources (AnGR) as applied throughout the report is an abbreviation of animal genetic resources used for food and agriculture and excludes fish.

xxvii

The length of the reporting process also means that the information presented in the SoWAnGR does not necessarily reflect the very latest developments in the state of institutions and capacity at the national level.

TABLE 1 Regional overview of Country Reports Region3

COUNTRY REPORTS Submitted Final

Total Draft

Africa

45

4

49

Asia

22

4

26

Europe and the Caucasus

38

3

41

Latin America and the Caribbean

21

9

30

Near and Middle East

6

3

9

North America

2

0

2

Southwest Pacific Total

9

3

12

143

26

169

Reports received by 31 December 2005.

3

Note that these regions do not correspond to the usual FAO regions; see below for further explanation.

xxviii

1

TABLE 2 Country Reports received Region

Countries

Africa (49)

Algeria, Angola, Benin, Botswana, Burkina Faso, Burundi, Cameroon, Cape Verde, Central African Republic, Chad, Comoros, Congo, Côte d’Ivoire, Democratic Republic of the Congo, Djibouti, Equatorial Guinea, Eritrea, Ethiopia, Gabon, Gambia, Ghana, Guinea, Guinea-Bissau, Kenya, Lesotho, Madagascar, Malawi, Mali, Mauritania, Mauritius, Morocco, Mozambique, Namibia, Niger, Nigeria, Rwanda, Sao Tome and Principe, Senegal, Seychelles, Sierra Leone, Somalia, South Africa, Swaziland, Togo, Tunisia, Uganda, United Republic of Tanzania, Zambia, Zimbabwe

Asia (26)

Afghanistan, Bangladesh, Bhutan, Cambodia, China, India, Indonesia, Iran (Islamic Republic of), Japan, Kazakhstan, Kyrgyzstan, Lao People’s Democratic Republic, Malaysia, Maldives, Mongolia, Myanmar, Nepal, Pakistan, Papua New Guinea, Philippines, Republic of Korea, Sri Lanka, Tajikistan, Turkmenistan, Uzbekistan, Viet Nam

Europe and the Caucasus (41)

Albania, Armenia, Austria, Azerbaijan, Belarus, Belgium, Bosnia and Herzegovina, Bulgaria, Croatia, Cyprus, Czech Republic, Denmark, Estonia, Finland, France, Georgia, Germany, Greece, Hungary, Iceland, Ireland, Italy, Latvia, Lithuania, Moldova, Netherlands, Norway, Poland, Portugal, Romania, Russian Federation, Serbia and Montenegro4, Slovakia, Slovenia, Spain, Sweden, Switzerland, The former Yugoslav Republic of Macedonia, Turkey, Ukraine, United Kingdom

Latin America and the Caribbean (30)

Antigua and Barbuda, Argentina, Barbados, Bolivia, Brazil, Chile, Colombia, Costa Rica, Cuba, Dominica, Dominican Republic, Ecuador, El Salvador, Grenada, Guatemala, Guyana, Haiti, Honduras, Jamaica, Mexico, Nicaragua, Panama, Paraguay, Peru, Saint Kitts and Nevis, Saint Lucia, Suriname, Trinidad and Tobago, Uruguay, Venezuela (Bolivarian Republic of)

Near and Middle East (9)

Egypt, Iraq, Jordan, Lebanon, Libyan Arab Jamahiriya, Oman, Sudan, Syrian Arab Republic, Yemen

North America (2)

Canada, United States of America

Southwest Pacific (12)

Australia, Cook Islands, Fiji, Kiribati, Niue, Northern Mariana Islands, Palau, Samoa, Solomon Islands, Tonga, Tuvalu, Vanuatu

Reports received by 31 December 2005.

Reports from international organizations Following a request from the ITWG, in August 2004, FAO invited 77 international organizations to submit a report of their work in the field of AnGR, as a contribution to the SoW-AnGR. These reports were to cover activities such as research, extension, education, training, public awareness, communications and advocacy, and also to include a description of the organization and information on institutional capacities which support activities in AnGR. Specific subjects to be described included (if applicable) inventory and characterization, sustainable use and development, conservation, valuation, policy and legislation, documentation and information databases, animal and human health, and food safety, as well as opportunities and proposals for interaction with other organizations and United Nations agencies. As of June 2006, nine organizations had submitted reports (Table 3). Reports were received from four international non-governmental organizations, three intergovernmental organizations, and two research organizations. A further three international organizations informed FAO that they were not engaged in AnGR-related activities. 4

Since June 2006 Serbia and Montenegro have become independent states. However, in the SoW-AnGR they are still treated as one country, as in Country Report submitted to FAO.

xxix

TABLE 3 Reports from international organizations Organization

Title of the submission

Received

CGIAR Centres

Consultative Group on International Agricultural Research (CGIAR) Centres Report to FAO for input into the SoW and the draft report on strategic priorities for action on FAnGR Section I: Description of the CGIAR Institutes and Programmes

May 2004

SAVE Foundation

SAVE Foundation (Safeguard for Agricultural Varieties in Europe) Brief Portrait April 2004

May 2004

D8 Countries

Report on Animal Genetic Resources in the D-8 Countries – Strategic Priorities for Action; and Reports D8 Seminar on Conservation of Farm Animal Genetic Resources Cairo, Egypt, 11–13 January 2004 D8 Seminar on Conservation of Farm Animal Genetic Resources, Islamabad, Pakistan, 1–3 August 2002; Report on Workshop on Food Security in D 8 countries, Babolsar, Islamic Republic of Iran, 16–20 October 2000 Report on Workshop on Food Security in D 8 countries, Islamabad, Pakistan, 24–26 November 1999

June 2004

LPP

League for Pastoral Peoples Report on Activities of the League for Pastoral Peoples

November 2004

OIE

World Organisation for Animal Health (OIE) Oral presentation to the Commission on Genetic Resources for Food and Agriculture, 10th Session (to be used thereafter as the OIE input in reply to the FAO AN21/47 request)

November 2004

ACSAD

Arab Center for the Studies of Arid zones and Dry lands (ACSAD) The Activities of the Arab Center for the Studies of Arid zones and Dry lands concerning the Animal Genetic Resources

December 2004

IAMZ

The Mediterranean Agronomic Institute of Zaragoza (IAMZ) Report on Training activities

January 2005

EAAP

EAAP (European Association for Animal Production) Report of the Working Group on Animal Genetic Resources

February 2005

ISAG

International Society for Animal Genetics (ISAG) Report of the ISAG/FAO advisory group on animal genetic diversity

March 2005

September 2004

Thematic studies In addition to the Country Reports and the reports from international organizations, a number of thematic studies were commissioned by FAO. These studies were intended to contribute to the understanding of specific topics likely not to be covered in Country Reports, but relevant to the preparation of the SoW-AnGR. During the period 2002 to 2006, 12 thematic studies were prepared: • Opportunities for incorporating genetic elements into the management of farm animal diseases: policy issues. A review paper on the potential of genetic elements in the management of disease, technical opportunities, and benefits arising from the incorporation of these elements in effective disease management5 (2002); • Measurement of domestic animal diversity (MoDAD) – a review of recent diversity studies. A survey evaluating the current status of molecular genetic research in domestic animal species, with emphasis on characterization of AnGR6 (2004); 5 6

Background Study Paper No. 18 CGRFA/WG-AnGR-3/04 inf. 3

xxx

• The economics of farm animal genetic resource conservation and sustainable use: why is it important and what have we learned? A study on the valuation of AnGR, summarizing methodological approaches and knowledge gaps7 (2004); • Conservation strategies for animal genetic resources. A study contrasting opportunities, challenges, biological characteristics, institutional infrastructure and operational considerations influencing management of plant and animal genetic resources8 (2004); • Environmental effects on animal genetic resources. An evaluation and synthesis of the evidence available on a spectrum of environmental factors and their effects on AnGR at the individual animal and the breeding population levels9 (2004); • The legal framework for the management of animal genetic resources. An introductory study of policy and legal frameworks for the management of AnGR including a survey of countries in different world regions10 (2004, printed revised version 2005); • The impact of disasters and emergencies on animal genetic resources. A study which provides an overview of potential disasters and their possible impact on AnGR. It also provides an analysis of the effects of emergency responses. It proposes decision-support guidelines for disaster management11 (2006); • The state of development of biotechnologies as they relate to the management of animal genetic resources and their potential application in developing countries. An introductory study of biotechnology applications and their use in developing countries, which includes information provided in Country Reports12 (2006); • Exchange, use and conservation of animal genetic resources: policy and regulatory options. A study which identifies how exchange practices related to AnGR affect the various stakeholders in the livestock sector (2006); • A strategic approach for conservation and continued use of farm animal genetic resources. A study which outlines patterns of change in AnGR use and their impact on conservation. It summarizes current experience, and the capacity of alternative conservation measures, considering the needs and aspirations of the various stakeholders whose livelihoods depend on animal production13 (2006); • People and animals. Traditional livestock keepers: guardians of domestic animal diversity. A documentation of 13 case studies from all over the world on how communities manage their local AnGR, demonstrating the value of local knowledge in preserving the equilibrium between farmers, animals and environment14 (2007); • Gene flow in animal genetic resources. A study on status, impact and trends. A study providing analysis of the magnitude and direction of movement of genetic material of the four major farm animal species: cattle, pigs, goats, and sheep. Determining factors are identified and selected; examples of impacts on economic development, poverty reduction and biodiversity in developing countries are presented (2007). 7

Background Study Paper No. 21 Background Study Paper No. 22 9 Background Study Paper No. 28 10 Background Study Paper No. 24 11 Background Study Paper No. 32 12 Background Study Paper No. 33 13 CGRFA/WG-AnGR-4/06/Inf.6 14 FAO Inter-Departmental Working Group on Biological Diversity for Food and Agriculture 8

xxxi

Preparation of the report Sources of information Different sections of the SoW-AnGR required different approaches. Some sections were largely based on the information provided in the 148 Country Reports available by June 2005. Other sections drew heavily on the wider literature or on expert knowledge rather than on the information gathered specifically for the SoW-AnGR process. FAO’s Domestic Animal Diversity Information System (DAD-IS)15 and the FAOSTAT16 statistical database were also utilized. Regional e-mail consultations, organized by FAO in late 2005 to review the draft Report on Strategic Priorities for Action, provided an additional source of information, particularly on institutional capacities. Part 1 describes the state of agricultural diversity in the livestock sector. The chapter draws on a number of sources. The description of AnGR inventory and of the extent of genetic erosion is based on information drawn from DAD-IS. This information system, which was launched in 1996, enables National Coordinators for the Management of Animal Genetic Resources to update their national breed databank via the Internet. The guidelines for the development of Country Reports encouraged countries to report breed-related data and information directly to DAD-IS, and not to include details of breeds in the Country Reports. Nonetheless, the Country Reports contained a wealth of breed-related information that was not reported to DAD-IS. As a result of this development, and in order to ensure that the analysis for the SoW-AnGR was based on the most up-to-date information available, FAO provided for the extraction of these data from Country Reports and their entry into DAD-IS. National Coordinators were then requested to validate and further complete their national breed databanks. It was also thought desirable to enable the analysis for the SoW-AnGR to be based on breeds and not only on national breed populations; i.e. that populations of the same breed in different countries were not counted as separate breeds. To this end, linkages between breed populations in different countries were introduced into the Global Databank, based on information on names, origin and development, importation and geographic location. Lists of all national breed populations and their proposed linkages were sent to National Coordinators for review. The analysis of the data for the purposes of the SoW-AnGR was carried out in January 2006, by which time data from all 169 Country Reports had been entered into the system. The section on uses and values of AnGR is based on FAOSTAT for population and production statistics, and on the Country Reports for more qualitative information on livestock functions. The section on genetic resistance to disease draws on DAD-IS and the wider scientific literature. Broader sources were also used to describe the origin and domestication of AnGR, sharing and exchange of AnGR, and threats to AnGR. Part 2 describes livestock sector trends and their implications for AnGR, and draws on a wide range of literature and statistics. Part 3 describes the state of human capacity, breeding and conservation strategies, legislation and the use of biotechnologies. This part of the report is largely based on the information in the Country Reports. However, the sections on regional and international legislation, and emerging legal and policy issues draw on wider sources. 15 16

http://www.fao.org/dad-is/ http://faostat.fao.org/

xxxii

Part 4 on the state of the art in AnGR management is largely based on the wider scientific literature. For the preparation of the section on the state of the art in AnGR conservation, an expert meeting was convened at FAO in Rome, in July 2005. The participants discussed the approach to the section and allocated writing tasks. The first draft was reviewed by all members in the writing group in October 2005. In November 2005, a workshop “Options and Strategies for the Conservation of Farm Animal Genetic Resources” took place in Montpellier, France. The participants at this workshop were given the opportunity to review the revised version of the conservation section. Part 5 analyses needs and challenges for AnGR management, based on the evidence provided in the other chapters of the report. This analysis relates the current state of erosion and threats to AnGR to current capacities in AnGR management and the state of knowledge regarding methodologies and their application.

Regional classification of countries The assignment of countries to the regions and subregions used for the purposes of the SoW-AnGR was based on a number of factors that influence biodiversity, including production environments, cultural specificities and the distribution of shared AnGR. Future collaboration in the establishment of Regional Focal Points was also considered, as was the experience gained from the process of convening SoW-AnGR subregional follow-up workshops in 2003 and 2004. Thus, the assignments do not follow exactly the standard FAO regions used in FAO statistics or for FAO election purposes (although for most countries the assignment does not differ from the standard classification). The proposed classification was reviewed at a meeting of Regional Facilitators on “Strategy for Regional Consultations” held in August 2005. The resulting classification distinguishes seven regions, of which three regions were further subdivided: Africa (East Africa, North and West Africa, Southern Africa); Asia (Central Asia, East Asia, Southeast Asia, South Asia); Europe and the Caucasus; Latin America and the Caribbean (Caribbean, Central America, South America); the Near and Middle East; North America; and the Southwest Pacific. FIGURE 1 Assignment of countries to regions and subregions in this report

State of the World : regions and subregions North America North America

Latin America and the Caribbean

Africa

Europe and the Caucasus

Caribbean

North and West Africa

South America Central America

Southern Africa East Africa

Europe and the Caucasus

Near and Middle East Near and Middle East

Asia

Southwest Pacific Central Asia

Southwest Pacific

South Asia East Asia Southeast Asia

xxxiii

Executive summary

T

he State of the World’s Animal Genetic Resources for Food and Agriculture is the first global assessment of livestock biodiversity. Drawing on 169 Country Reports, contributions from a number of international organizations and twelve specially commissioned thematic studies, it presents an analysis of the state of agricultural biodiversity in the livestock sector – origins and development, uses and values, distribution and exchange, risk status and threats – and of capacity to manage these resources – institutions, policies and legal frameworks, structured breeding activities and conservation programmes. Needs and challenges are assessed in the context of the forces driving change in livestock production systems. Tools and methods to enhance the use and development of animal genetic resources are explored in sections on the state of the art in characterization, genetic improvement, economic evaluation and conservation. Thousands of years of animal husbandry and controlled breeding, combined with the effects of natural selection, have given rise to great genetic diversity among the world’s livestock populations. High-output animals – intensively bred to supply uniform products under controlled management conditions – co-exist with the multipurpose breeds kept by small-scale farmers and herders mainly in low external input production systems. Effective management of animal genetic diversity is essential to global food security, sustainable development and the livelihoods of hundreds of millions of people. The livestock sector and the international community are facing many challenges. The rapidly rising demand for livestock products in many parts of the developing world, emerging animal diseases, climate change and global targets such as the Millennium Development Goals need to be urgently addressed. Many breeds have unique characteristics or combinations of characteristics – disease resistance, tolerance of climatic extremes or supply of specialized products – that could contribute to meeting these challenges. However, evidence suggests that there is ongoing and probably accelerating erosion of the genetic resource base. FAO’s Global Databank for Animal Genetic Resources for Food and Agriculture contains information on a total of 7 616 livestock breeds. Around 20 percent of reported breeds are classified as at risk. Of even greater concern is that during the last six years 62 breeds became extinct – amounting to the loss of almost one breed per month. These figures present only a partial picture of genetic erosion. Breed inventories, and particularly surveys of population size and structure at breed level, are inadequate in many parts of the world. Population data are unavailable for 36 percent of all breeds. Moreover, among many of the most widely used high-output breeds of cattle, within-breed genetic diversity is being undermined by the use of few highly popular sires for breeding purposes. A number of threats to genetic diversity can be identified. Probably the most significant is the marginalization of traditional production systems and the associated local breeds, driven mainly by the rapid spread of intensive livestock production, often large-scale and utilizing a narrow range of breeds. Global production of meat, milk and eggs is increasingly based on a limited number of high-output breeds – those that are most profitably utilized in industrial production systems. The intensification process has been driven by rising demand for animal products and has been facilitated by the ease with which genetic material, production technologies and inputs can now be moved around the world. Intensification and industrialization have contributed to raising the output of livestock production and to feeding the growing human population. However, policy measures are necessary to minimize the potential loss of the global public goods embodied in animal genetic resource diversity.

xxxv

Acute threats such as major disease epidemics and disasters of various kinds (droughts, floods, military conflicts, etc.) are also a concern – particularly in the case of small, geographically concentrated breed populations. Threats of this kind cannot be eliminated, but their impacts can be mitigated. Preparedness is essential in this context as ad hoc actions taken in an emergency situation will usually be far less effective. Fundamental to such plans, and more broadly to the sustainable management of genetic resources, is improved knowledge of which breeds have characteristics that make them priorities for conservation, and how they are distributed geographically and by production system. Policies and legal frameworks influencing the livestock sector are not always favourable to the sustainable utilization of animal genetic resources. Overt or hidden governmental subsidies have often promoted the development of large-scale production at the expense of the smallholder systems that utilize local genetic resources. Development interventions and disease control strategies can also pose a threat to genetic diversity. Development and post-disaster rehabilitation programmes that involve livestock should assess their potential impacts on genetic diversity and ensure that the breeds used are appropriate to local production environments and the needs of the intended beneficiaries. Culling programmes implemented in response to disease outbreaks need to incorporate measures to protect rare breeds; revision of relevant legislation may be necessary. Where the evolution of livestock production systems threatens the ongoing use of potentially valuable genetic resources, or to safeguard against sudden disastrous losses, breed conservation measures have to be considered. In vivo conservation options include dedicated conservation farms or protected areas, and payments or other support measures for those who keep rare breeds within their production environments. In vitro conservation of genetic material in liquid nitrogen can provide a valuable complement to in vivo approaches. Where feasible, facilitating the emergence of new patterns of sustainable utilization should be an objective. Particularly in developed countries, niche markets for specialized products, and the use of grazing animals for nature or landscape management purposes, provide valuable opportunities. Well-planned genetic improvement programmes will often be essential if local breeds are to remain viable livelihood options for their keepers. Implementing appropriate strategies for the low external input production systems of the developing world is a great challenge. Pastoralists and smallholders are the guardians of much of the world’s livestock biodiversity. Their capacity to continue this role may need to be supported – for example by ensuring sufficient access to grazing land. At the same time, it is essential that conservation measures do not constrain the development of production systems or limit livelihood opportunities. A small number of community-based conservation and breeding programmes have begun to address these issues. The approach needs to be further developed. Effective management of animal genetic diversity requires resources – including welltrained personnel and adequate technical facilities. Sound organizational structures (e.g. for animal recording and genetic evaluation) and wide stakeholder (particularly breeders and livestock keepers) involvement in planning and decision-making are also essential. However, throughout much of the developing world, these prerequisites are lacking. Forty-eight percent of the world’s countries report no national-level in vivo conservation programmes, and sixty-three percent report that they have no in vitro programmes. Similarly, in many countries structured breeding programmes are absent or ineffective.

xxxvi

In a time of rapid change and widespread privatization, national planning is needed to ensure the long-term supply of public goods. Livestock-sector development policies should support equity objectives for rural populations so that these populations are able to build up, in a sustainable way, the productive capacity required to enhance their livelihoods and supply the goods and services needed by the wider society. The management of animal genetic resources needs to be balanced with other goals within the broader rural and agricultural development framework. Careful attention must be paid to the roles, functions and values of local breeds and to how they can contribute to development objectives. The countries and regions of the world are interdependent in the utilization of animal genetic resources. This is clear from evidence of historic gene flows and current patterns of livestock distribution. In the future, genetic resources from any part of the world may prove vital to breeders and livestock keepers elsewhere. There is a need for the international community to accept responsibility for the management of these shared resources. Support for developing countries and countries with economies in transition to characterize, conserve and utilize their livestock breeds may be necessary. Wide access to animal genetic resources – for farmers, herders, breeders and researchers – is essential to sustainable use and development. Frameworks for wide access, and for equitable sharing of the benefits derived from the use of animal genetic resources, need to be put in place at both national and international levels. It is important that the distinct characteristics of agricultural biodiversity – created largely through human intervention and requiring continuous active human management – are taken into account in the development of such frameworks. International cooperation, and better integration of animal genetic resources management into all aspects of livestock development, will help to ensure that the world’s wealth of livestock biodiversity is suitably used and developed for food and agriculture, and remains available for future generations.

xxxvii

PART 1

Introduction

The importance of the world’s biodiversity – the variety of its plants, animals and microorganisms, and of the ecosystems of which they form a part, is increasingly recognized. Agricultural biodiversity encompasses the diversity of the cultivated plants and domestic animals utilized by humankind for the production of food and other goods and services. More broadly, it includes the diversity of the agro-ecosystems on which this production depends. The capacity of agro-ecosystems to maintain and increase their productivity, and to adapt to changing circumstances, is vital to the food security of the world’s population. The 40-plus livestock species contributing to today’s agriculture and food production are shaped by a long history of domestication and development. Selection pressures resulting from environmental stress factors, and the controlled breeding and husbandry imposed by humans, have combined to produce a great variety of genetically distinct breeds1. This diversity, developed over thousands of years, is a valuable resource for today’s livestock keepers. Genetically diverse livestock populations provide a greater range of options for meeting future challenges, whether associated with environmental change, emerging disease threats, new knowledge of human nutritional requirements, fluctuating market conditions or changing societal needs. Part 1 of the Report begins by describing the origin of the diversity of today’s animal genetic resources for food and agriculture (AnGR) – the domestication and history of livestock species. This is followed by a description of the current status of AnGR diversity on a global scale, and the extent to which this diversity is threatened by genetic erosion. The next section describes patterns of international exchange of AnGR. The roles and values of AnGR, and their direct and indirect contributions to livelihoods and economic output in the various regions of the world are then outlined. The importance of genetic resistance to disease as a resource in the field of animal health is also introduced. In the final section of Part 1, threats to the world’s AnGR diversity are discussed.

1

Central to the description of livestock diversity is the notion of the breed (see Part 4 – Section A: 1 for a discussion of the definition of the term “breed”)

THE STATE OF AG R I CULT U R AL B I OD I VERSI T Y I N THE L I VESTOC K SEC TO R

Section A

Origin and history of livestock diversity 1

Introduction

The history of AnGR started around 12 000 to 14 000 years ago, during the agricultural revolution of the early Neolithic, with the domestication of major crop and livestock species. This control of food production led to major demographic, technological, political and military changes. The domestication of animals and plants is considered to be one of most important developments in history, and one of the prerequisites for the rise of human civilizations (Diamond, 2002). After the initial domestication events, the spread of farming into nearly all terrestrial habitats followed rapidly (Diamond and Bellwood, 2003; Figure 2).

Subsequently, thousands of years of natural and human selection, genetic drift, inbreeding and cross-breeding have contributed to AnGR diversity and have allowed livestock keeping to be practised in a variety of environments and production systems. AnGR diversity is vital to all production systems. It provides the raw material for breed improvement, and for adaptation to changing circumstances. As revealed by recent molecular studies, the diversity found in today’s indigenous livestock populations and breeds greatly exceeds that found in their commercial counterparts. Unravelling the origin

FIGURE 2 Archaeological map of agricultural homelands and spread of Neolithic/Formative cultures, with approximate radiocarbon dates

Fertile Crescent 11000 BP Eastern USA 4000 - 3000 BP

Yangzi & Yellow River Basins 9000 BP

central Mexico

5000 - 4000 BP

Amazonia?

Sub-Saharan Africa? 5000 - 4000 BP New Guinea highlands 9000 - 6000 BP

northern South America

Approximate limits of prehistoric agriculture (deserts, mountains etc. not differentiated)

Map drawn by Clive Hilliker and provided by Peter Bellwood.

5

THE STATE OF THE WO RLD' S A N I MAL GENE T I C RESOU RC ES FO R FOOD A ND AG R I CULT U RE

PART 1

and distribution of livestock diversity is central to its current utilization, and to its long-term conservation (Hanotte et al., 2006).

2

The livestock domestication process

Very few animal species have been successfully domesticated. Domestication was a complex and gradual process, which altered the behaviour and morphological characteristics of the ancestral animals (Box 1). The circumstances and pressures that triggered the domestication of animals remain uncertain, and may have varied from one geographic area to another and from one species to another. The roots of animal domestication are probably related to the ubiquitous tendency of hunter gatherers (presumably shared by early humans) to try to tame or manage wild animals (Diamond, 2002). It was, however, at the end of the Pleistocene that the process of domestication actually got underway. At this time, changes in the climate, which became more unpredictable, warmer and/or more seasonal in some areas, led to localized expansion of human populations. These developments triggered the uptake of crop farming, and affected the distribution and density of the wild species hunted for food. In these circumstances, the main driver of animal domestication may have been the desire to secure

the availability of “favourite” foods – with the potential of some domesticated species to provide support to crop farming (e.g. ploughing with oxen or buffalo), or as pack and riding animals (e.g. llamas, dromedaries, Bactrian camels, horses, donkeys and even cattle) being realized later. Among the world’s 148 non-carnivorous species weighing more than 45 kg, only 15 have been domesticated. Thirteen of these species are from Europe and Asia, and two originate from South America. Moreover, only six have become widespread on all continents (cattle, sheep, goats, pigs, horses, and donkeys), while the remaining nine (dromedaries, Bactrian camels, llamas, alpacas, reindeer, water buffalo, yaks, Bali cattle, and mithun) are important in more limited areas of the globe (adapted from Diamond, 1999). The proportion is even lower in the case of birds, with only ten species (chickens, domestic ducks, Muscovy ducks, domestic geese, guinea fowl, ostriches, pigeons, quails, and turkeys) currently domesticated out of around 10 000 avian species (the list excludes the many birds domesticated for ornamental or recreational purposes). With the exception of the wild boar (Sus scrofa) the ancestors and wild relatives of major livestock species are either extinct or highly endangered as a result of hunting, changes to their habitats, and in the case of the wild red jungle fowl, intensive cross-breeding with the domestic counterpart. In these species, domestic livestock are the only depositories of the now largely vanished diversity

Box 1 The domestication process Domesticated animals are here considered to be those species that are bred in captivity, and modified from their wild ancestors to make them more useful to humans, who control their reproduction (breeding), care (shelter, protection against predators) and food supply (Diamond, 2002; Mignon-Grasteau, 2005). Domestication includes the following steps: initial association with free breeding; confinement; confinement with breeding in captivity; and selective

6

breeding and breed improvement (modified from Zeuner, 1963). Archaeologists and animal geneticists use various means to unravel the history of domestication, including study of morphological changes to the teeth, cranium and skeleton; and the construction of demographic age and sex curves which allow the identification of patterns indicative of domestication (Zeder et al., 2006).

THE STATE OF AG R I CULT U R AL B I OD I VERSI T Y I N THE L I VESTOC K SEC TO R

TABLE 4 Origin and domestication of livestock species Domestic species

Cattle

Bos taurus taurus

Wild Ancestor

MtDNA

Domestication

Time

clades

events*

B.P.

Location

Aurochs (3 subspecies) (extinct)

B. primigenius primigenius

4

1

~ 8000

Near & Middle East (west Asia)

B. p. opisthonomous

2

1

~ 9500

northeast Africa

B. p. nomadicus

2

1

~ 7000

northern Indian subcontinent

3

1

~ 4500

Qinghai-Tibetan Plateau

5

2

~ 10000

Near and Middle East, northern Indian subcontinent

4

2

~ 8500

Near and Middle East/Turkey (Central Anatolia)

Riverine B. bubalus bubalus

ND

1

~ 5000

Islamic Republic of Iran/Iraq, Indian subcontinent

Swamp B. bubalus carabensis

ND

1

~ 4000

Southeast Asia, China

6

6

~ 9000

Europe, Near and Middle East, China

Bos taurus indicus Yak

Poephagus grunniens Goat

Capra ferus

Sheep

Ovis aries

Water buffalo

Pig

Sus scrofa domesticus

Wild yak

P. mutus Bezoar Capra aegragus (3 subspecies)

Asian mouflon

Ovis orientalis

Asian wild buffalo

Wild boar

Sus scrofa (16 subspecies)

Indian subcontinent, Southeast Asia Horse

Extinct

Equus caballus Donkey

Equus asinus

17

multiple

~ 6500

Eurasian steppe

~ 6000

northeast Africa

~ 6500

Andes

African wild donkey

Equus africanus Nubian wild ass E. a. africanus

1

1

Somali wild ass E. a. somali

1

1

ND

1

Llama

Lama glama

2 subspecies

L. guanicoe guanicoe L. guanicoe cacsiliensis

7

THE STATE OF THE WO RLD' S A N I MAL GENE T I C RESOU RC ES FO R FOOD A ND AG R I CULT U RE

PART 1

TABLE 4 cont. Origin and domestication of livestock species Domestic species

Wild Ancestor

MtDNA

Domestication

Time

clades

events*

B.P.

Location

Alpaca

Vicugna pacos

2 subspecies

ND

1

~ 6500

Andes

ND

1

~ 4500

Central Asia (eastern Islamic Republic of Iran)

ND

1

~ 5000

southern Arabian Peninsula

5

2

V. vicugna vicugna V. vicugna mensalis Bactrian Camel

Extinct**

Camelus bactrianus C. b. ferus

Dromedary

Extinct

Camelus dromedarius Domestic chicken

Gallus domesticus

Red Junglefowl

Gallus gallus (4 subspecies G. g. spadiceus, G. g. jabouillei

~ 5000

Indian subcontinent

~ 7500

China – Southeast Asia

G. g. murghi, G. g. gallus) Source: adapted and updated from Bruford et al. (2003); FAO (2005). *Minimum number of domestication events.**Recent genetic evidence suggests that the endangered wild population are not the ancestral maternal populations of today’s domestic Bactrian (Jianlin et al., 1999). ND = not determined.

of the wild ancestors (Table 4). This is a major difference from crop species, in many of which the wild ancestors are commonly found at the centres of origin and represent an important source of variation and adaptive traits for future breeding programmes. The small number of animal species successfully domesticated is largely explained by the characteristics required (or advantageous) for domestication, which are rarely found together in a single species. All major livestock species were domesticated several thousand years ago. It is improbable that further large mammalian species will be domesticated, at least in the near future, as illustrated by the failure, or at best only partial success, of twentieth century

8

attempts to domesticate new species (e.g. oryx, zebras, African buffaloes and various species of deer). However, the coming years may see further development of the captive breeding of small and “non-conventional” species (sometimes called microlivestock) for human consumption, which may become more important, at least locally or regionally (BOSTID, 1991; Hanotte and Mensah, 2002). Important or essential characteristics for successful domestication include behavioural traits such as a lack of aggression towards humans; a strong gregarious instinct, including “follow the leader” dominance hierarchies which allow the possibility of a human substitute as leader; a tendency not to panic when disturbed;

THE STATE OF AG R I CULT U R AL B I OD I VERSI T Y I N THE L I VESTOC K SEC TO R

the ability to breed in captivity; physiological traits such as a diet that can easily be supplied by humans (domestication of herbivores rather than carnivores); a rapid growth rate; relatively short intervals between births; and large litter size (Diamond, 2002). The ancestral species of the majority of livestock species have now been identified (Table 4). It is also known that many current domestic animal populations and breeds originate from more than one wild ancestral population, and that in some cases there has been genetic admixture or introgression between species that do not normally hybridize in the wild. These admixture and hybridization events probably occurred after

the initial domestication. They were often linked to human migration, trading or simply the result of the requirement of agricultural societies for new livestock phenotypes. Examples include admixture between taurine and Zebu cattle, the presence of cattle genetic background in yaks and Bali cattle, Asian pig hybridization with European breeds, cross-breeding between dromedaries and Bactrian camels, and (as revealed by recent genetic studies) intensive admixture between the two South American domestic camelids (llamas and alpacas) (Kadwell et al., 2001).

Box 2 Molecular characterization – a tool to understand livestock origin and diversity Recent major developments in molecular genetics have provided powerful new tools, called molecular markers, to assess the origins of livestock species and the geographic distribution of their diversity. Protein polymorphisms were the first molecular markers used in livestock. A large number of studies, particularly during the 1970s, documented the characterization of blood group and allozyme systems. However, the level of polymorphism observed in proteins is often low, which reduces the general applicability of protein typing in diversity studies. DNA-based polymorphisms are now the markers of choice for molecular-based surveys of genetic diversity. Importantly, polymorphic DNA markers showing different patterns of Mendelian inheritance can be studied in nearly all major livestock species. Typically, they include D-loop and cytochrome B mitochondrial DNA (mtDNA) sequences (maternal inheritance), Y chromosome-specific single nucleotide polymorphisms (SNPs) and microsatellites (paternal inheritance), and autosomal microsatellites (biparental inheritance). Autosomal microsatellites have been isolated in large numbers from most livestock species. FAO/ISAG (International Society of

Animal Genetics) recommended lists of autosomal microsatellite markers for genetic diversity studies are publicly available (http://www.fao.org/dad-is). Different genetic markers provide different levels of genetic diversity information. Autosomal microsatellite loci are commonly used for population diversity estimations, differentiation of populations, calculation of genetic distances, estimation of genetic relationships, and the estimation of population genetic admixture. MtDNA sequences are the markers of choice for domestication studies, as the segregation of an mtDNA lineage within a livestock population will only have occurred through the domestication of a wild female, or through the incorporation of a female into the domestic stock. More particularly, mtDNA sequences are used to identify putative wild progenitors, the number of maternal lineages and their geographic origins. Finally, the study of a diagnostic Y chromosome polymorphism is an easy and rapid way to detect and to quantify malemediated admixture. Reproduced and adapted from FAO (2005).

9

THE STATE OF THE WO RLD' S A N I MAL GENE T I C RESOU RC ES FO R FOOD A ND AG R I CULT U RE

PART 1

3

Ancestors and geographic origins of our livestock

One of the most exciting areas of intersection between archaeology and genetics has been in documenting the locations of livestock domestication (Zeder et al., 2006), with archaeology guiding genetic research, and genetics providing support to some controversial archaeological theories or revealing possible new geographic origins for livestock species and their diversity. More particularly, it is now known that nearly all major livestock species are the result of multiple domestication events in distinct geographic areas (Table 4 and Figure 3); and that subsequent to the initial domestication events, genetic introgression between wild relatives and their domestic counterparts often occurred. It should be noted that apparently independent livestock domestication events were not necessarily culturally independent. Some independent domestication events may have represented the movement of a few domesticated individuals into a new area, with the genetic signatures of the

introduced founders subsequently submerged by the recruitment of local wild animals (Zeder et al., 2006). Alternatively, ancient signatures of local domestication events may now be hidden by more recent arrivals of livestock from other centres of origin. Osteometric information from archaeological sites, and ancient livestock DNA studies are important tools to address these questions. Livestock domestication is now thought to have occurred in at least 12 areas of the world (Figure 3). Interestingly, not all centres of domestication are closely associated with the homelands of our crop species (see Figure 2). While in some cases (e.g. the Fertile Crescent), domestication centres of both crops and livestock are intermingled, in others (e.g. the African continent) crop and livestock domestication seem largely to have occurred independently. While uncertainties still surround the existence of some domestication centres for some species, the following geographic areas are important primary centres of origin, and therefore diversity, of livestock species: the Andean chain of South America (llamas, alpacas, guinea pigs); central America

FIGURE 3 Major centres of livestock domestication – based on archaeological and molecular genetic information

7 3

12

5 8

1

4

11

9

6 10

2

(1) turkey (2) guinea pig, llama, alpaca, (3) pig, rabbit (4) cattle, donkey, (5) cattle, pig, goat, sheep, Bactrian camel (6) cattle, goat , chicken, river buffalo, (7) horse, (8) yak, (9) pig, swamp buffalo, chicken, (10) chicken, pig, Bali cattle (11) dromedary, (12) reindeer.

10

THE STATE OF AG R I CULT U R AL B I OD I VERSI T Y I N THE L I VESTOC K SEC TO R

(turkeys, Muscovy ducks); northeast Africa (cattle, donkeys); southwest Asia including the Fertile Crescent (cattle, sheep, goats, pigs); the Indus valley region (cattle, goats, chickens, riverine buffaloes); Southeast Asia (chickens, Bali cattle); east China (pigs, chicken, swamp buffaloes); the Himalayan plateau (yaks); and north Asia (reindeer). Additionally, the southern part of the Arabian Peninsula is thought to be the region of origin of the dromedary, the Bactrian camel may originate from the area that is now the Islamic Republic of Iran, and the horse from the Eurasian steppes. While domestication occurred in several places, it also happened at different times. Exact dating of domestication events has, however, proved particularly challenging. Animals undergoing the initial process of domestication would not have been significantly different in morphology from their wild ancestors, and dates relying on morphological markers will undoubtedly underestimate the age of domestication events (Dobney and Larson, 2006). The process of molecular dating, while independent of morphological changes, is typically characterized by large error rates, and often relies on uncertain calibration points. Approaches including demographic profiling techniques for identifying initial attempts at livestock management by humans, and calibration of molecular clocks using ancient DNA information, are providing new avenues for pinpointing the dates of domestication (Zeder et al., 2006). New archaeological and genetic information is constantly improving our understanding of the origin of livestock species. The first animal to be domesticated was the dog. This probably occurred at least 14 000 years ago – the animals being used for hunting and as watchdogs. It is unclear where the initial domestication took place, but many maternal lineages have been found in modern dogs – indicating multiple introgressions from their wild ancestor the grey wolf (Canis lupus) in the Old World. Domestic dogs were, apparently, not independently domesticated in the New World; the mitochondrial lineages identified so

far in the Americas are of European origin (Wayne et al., 2006). Goats were domesticated as early as 10 000 years ago in the Zagros Mountains of the Fertile Crescent (Zeder and Hesse, 2000). The bezoar (Capra aegragus) was probably one of the ancestors of the domestic goat, but it is possible that other species such as C. falconeri, contributed to the genetic pool of the domestic species. Today, five distinct maternal mitochondrial major lineages have been identified in domestic goats (Luikart et al., 2001; Sultana et al., 2003; Joshi et al., 2004). One of these lineages predominates numerically, and is present worldwide, while a second seems to be of contemporary origin. They probably reflect the primary caprine domestication process in the Fertile Crescent, where archaeological information suggests two to three areas of domestication (Zagros Mountains, Taurus Mountains, Jordan Valley). The other lineages are more restricted in their geographic distribution, and may correspond to additional domestications or introgressions in other areas including the Indus Valley (Fernández et al., 2006). Sheep were also probably first domesticated in the Fertile Crescent, approximately 8 000 to 9 000 years ago. Archaeological information suggests two independent areas of sheep domestication in Turkey – the upper Euphrates valley in eastern Turkey, and central Anatolia (Peters et al., 1999). Three species of wild sheep (the urial, Ovis vignei; the argali, O. ammon; and the Eurasian mouflon, O. musinom/orientalis) have been proposed as ancestors of domestic sheep (Ryder, 1984) or at least to have introgressed some local breeds. However, recent genetic work has indicated no contribution from the urial or argali (Hiendleder et al., 1998). This supports the view that the Asian mouflon (O. orientalis), which is found in a wide region stretching from Turkey at least as far as the Islamic Republic of Iran, is the only progenitor of domestic sheep. The European mouflon (O. musinom) is now considered to be a descendant of feral sheep. Four major maternal mitochondrial DNA lineages have been recorded in domestic sheep (Hiendleder et al., 1998; Pedrosa et

11

THE STATE OF THE WO RLD' S A N I MAL GENE T I C RESOU RC ES FO R FOOD A ND AG R I CULT U RE

PART 1

al., 2005; Tapio et al., 2006), one or two of which could correspond to distinct domestication events, and the others to subsequent wild introgression. To date, no clear associations have been described between these mitochondrial DNA lineages and phenotypic sheep varieties (e.g. fat-tailed, thintailed or fat-rumped sheep). The ancestor of the domestic pig is the wild boar (Sus scrofa). Extensive zooarchaeological findings indicate that pigs were domesticated around 9 000 years ago in the Near East. Material from several sites in eastern Anatolia indicates gradual changes in pig morphology and demographic profiles over several thousand years, providing evidence of the domestication process and its morphological consequences. Both archaeological and genetic evidence indicate a second major independent domestication centre in East Asia (China) (Guiffra et al., 2000). At least 16 distinct subspecies of wild boar have been described in Eurasia and North Africa and, perhaps not surprisingly, a recent survey of mitochondrial DNA diversity among Eurasian domestic pigs and wild boar revealed a complex picture of pig domestication, with at least five or six distinct centres across the geographic range of the wild species (Larson et al., 2005). Domestication of cattle has been particularly well documented, with clear evidence of three distinct initial domestication events for three distinct aurochs (Bos primigenius) subspecies. B. primigenius primigenius, domesticated in the Fertile Crescent around 8 000 years ago, and B. p. opisthonomous, possibly domesticated as early as 9 000 years ago in the northeastern part of the African continent (Wendorf and Schild, 1994), are the ancestors of the humpless B. taurus cattle of the Near East and Africa respectively. Humped Zebu cattle (Bos indicus), are now believed to have been domesticated at a later date, around 7 000 to 8 000 years ago, in the Indus Valley region of modern-day Pakistan (Loftus et al., 1994; Bradley et al., 1996; Bradley and Magee, 2006). Recently, a fourth domestication centre has been suggested in East Asia (Mannen et al., 2004), but it is unclear whether it occurred independently or represents local aurochs introgression in cattle of Near Eastern origin.

12

The ancestor of the domestic water buffalo (Bubalus bubalus) is undoubtedly the wild buffalo of Asia. Two main types are recognized, based on their phenotypes, karyotypes and recent mitochondrial DNA work (Tanaka et al., 1996): the riverine buffalo, found in the Indian subcontinent, the Near and Middle East, and eastern Europe; and the swamp buffalo, found in China and Southeast Asian countries. The two types hybridize in the northeastern part of the Indian subcontinent. They were probably domesticated separately, with possible centres of domestication of the riverine buffalo in the Indus Valley and/or the Euphrates and Tigris valleys some 5 000 years ago; and of the swamp buffalo in China, where it was domesticated at least 4 000 years ago in association with the emergence of rice cultivation. There is an ongoing debate as to when and where the horse (Equus caballus) was domesticated. The ancestor of the domestic horse is extinct. Two species have been regarded as putative wild ancestors – the tarpan (E. ferus) and the Przewalski horse (E. przewalskii). The Przewalski horse, although very closely related to the wild ancestor, is probably not the direct progenitor of the domestic species (Olsen et al., 2006; Vilà et al., 2006). It is difficult to assess whether archaeological horse remains are wild or domestic. Substantial evidence from north Kazakhstan (Botai culture) supports the view that horses were domesticated in this area during the Copper Age around 3700 to 3100 BC (Olsen, 2006). Recent molecular studies indicate that the diversity of the horse on the maternal side probably originates from several populations in different geographic areas. However, the data are not yet conclusive as to whether there was a single domestication event and subsequent introgression, or multiple independent domestication events (Vilà et al., 2001; Jansen et al., 2002). In contrast, the domestication of the donkey Equus asinus seems to have followed a much simpler process. Mitochondrial DNA studies have confirmed an African origin for the domestic donkey, and have ruled out the Asiatic wild ass as a possible progenitor (Beja-Pereira et al., 2004). Two mitochondrial lineages suggest

THE STATE OF AG R I CULT U R AL B I OD I VERSI T Y I N THE L I VESTOC K SEC TO R

two domestication events. One lineage is closely linked to the Nubian wild ass (E. asinus africanus), which is still found today living wild in northeastern Sudan close to the Red Sea. The other lineage shows some affinities to the Somali wild ass (E. asinus somaliensis). It could, therefore, also have an African origin, although domestication in a neighbouring area (Arabian Peninsula or Fertile Crescent) cannot be excluded. Archaeological evidence from Egypt supports an African centre of domestication for the donkey, and suggests a domestication date of around 6 000 to 6 500 years ago (Clutton-Brock, 1999). The domestic yak (Poephagus grunniens) is endemic to Central Asia and well adapted to a cold and high-altitude environment. Yak pastoralism is widespread in the Central Asian Highlands, and the introduction of yak pastoralism was crucial to the development of year-round sustainable occupation of the higher altitude zones of the Himalaya Plateau. It may have been connected with the establishment of Tibetan–Burman populations in this region. Today, some wild yaks (P. mutus) are still found on the Qinghai-Tibetan Plateau, but they may have been heavily introgressed with feral domestic yak. Three mitochondrial DNA lineages have been identified. However, similar geographic distributions of mitochondrial DNA diversity suggest a single domestication event in the eastern part of the Qinghai-Tibetan Plateau rather than multiple domestication events (Qi, 2004; Guo et al., 2006). Molecular findings also indicate that the dispersal of domestic yaks followed two separate migratory routes from their centre of domestication: the yak reached the “Pamir Knot” by following a westward route through the Himalaya and Kunlun Mountains; and reached Mongolia, and what is now the Russian Federation, by following a northward route through the Mongolian South Gobi and Gobi Altai Mountains (Qi et al., in press). As in the case of the yak, the domestication of the reindeer (Rangifer tarandus) has allowed pastoral communities to occupy habitats that would otherwise be largely unsuitable for livestock keeping. Very little is known about reindeer

domestication. The wild reindeer was possibly the latest large mammalian species to be domesticated. The oldest definitive archaeological evidence of reindeer domestication was discovered in the Altai Mountains of Siberia, and has been dated to about 2 500 years ago; it indicates that reindeer riding was practised at the time (Skjenneberg, 1984). There is no reliable information as to how reindeer domestication reached Europe; it could have developed independently in Scandinavia, or may have been adopted by the Saami people through contact with other north Eurasian pastoral communities. Reindeer husbandry is believed to have developed among the Saami sometime after 1600 AD. The wild reindeer is known as the caribou in North America; it is believed never to have been domesticated on this continent (Clutton-Brock, 1999). The domestication of the Bactrian camel (Camelus bactrianus) may have occurred in the area that is now the Islamic Republic of Iran/Turkmenistan, or further east, in southern Kazakhstan, northwestern Mongolia or northern China (Bulliet, 1975; Peters and von den Driesch, 1997). The earliest evidence of domestic Bactrian camels is from the site of Sahr-i Sokta in the central part of the Islamic Republic of Iran, from where camel bones, dung, and woven fibres dating from approximately 2600 BC have been recovered (Compagnoni and Tosi, 1978). Recent genetic work indicates that the wild camel (C. ferus) populations of the Gobi Desert, which successfully hybridize with the domestic species, are probably not the direct maternal ancestors of domestic or feral camels (Jianlin, et al., 1999). The wild ancestor of the one-humped dromedary (C. dromedarius) is now extinct. Domestication of the species is believed to have started around 5 000 years ago in the southeastern part of the Arabian Peninsula. The origin of the South American camelidae has now been unravelled, with the guanaco (Lama guanicoe) and the vicuña (Vicugna vicugna) being the ancestral species of the domestic llama (Lama glama) and alpaca (Vicugna pacos), respectively (Kadwell et al., 2001). Archaeozoological evidence

13

THE STATE OF THE WO RLD' S A N I MAL GENE T I C RESOU RC ES FO R FOOD A ND AG R I CULT U RE

PART 1

points to the central Peruvian Andes as the centre of origin of the alpaca, 6 000 to 7 000 years before present. The llama was probably domesticated at the same period in the Andes around Lake Titicaca. Large-scale introgressions between the two domestic species have been revealed (Wheeler et al., 2006) – an ongoing hybridization process which probably began with the Spanish conquest, which destroyed the traditional breeding structures and management of the two species. The ancestor of Bali cattle is the banteng (Bos javanicus), of which three endangered subspecies have been recognized. The domestication of the species did not, in fact, occur on the Island of Bali, where there is no evidence for the presence of the wild ancestor. The species could have been domesticated in Java and/or on the Indo–Chinese Peninsula. B. taurus and B. indicus introgression has been found in Bali cattle, and Bali cattle genetic background has also been inferred in several Southeast Asian cattle breeds, suggesting that the domestic species once had a wider distribution than it has today (Felius, 1995). The ancestor of the mithun (B. frontalis) is the gaur (B. gaurus). As in the case of Bali cattle, the centre of domestication of the species in unknown. Archaeological excavation in northeastern Thailand (Non Nok Tha) suggests that both species might have been domesticated as early as 7 000 years ago (Higham (1975) in Felius, 1995). The domestic chicken (Gallus domesticus) is descended from the wild red jungle fowl (Gallus gallus), with five possible progenitor subspecies. While previous molecular studies suggested a single domestic origin in Southeast Asia (Thailand) (Fumihito et al., 1994; 1996), at least six distinct maternal genetic lineages have now been identified (Liu et al., 2006), suggesting more than one domestication centre. Archaeological information indicates a centre of chicken domestication around the Indus Valley 5 000 years ago, and another in eastern China maybe as early as 7 500 to 8 000 years ago (West and Zhou, 1988).

14

4

Dispersal of domesticated animals

If the domestication process was the major initiating event in the development of today’s livestock diversity, the subsequent dispersion and migration of domesticated species across all five continents was equally important. This process played a major role in the emergence of the current geographic distribution of livestock diversity. The main factors at the root of the early dispersion of livestock species were the expansion of agriculture, trade and military conquests. The exact mechanisms through which agricultural expansion occurred are still debated. The process probably varied from one region to another (Diamond and Bellwood, 2003). It certainly involved both the movement of human populations, and cultural exchanges between populations – as illustrated by the adoption of farming by many hunter–gatherer societies. Important examples of agricultural expansions include that of the Neolithic, which brought cattle, sheep and goats into Europe, and may have triggered the local domestication of the wild boar. Domesticated livestock followed two distinct major routes into Europe – the Danubian and the Mediterranean (Bogucki, 1996; Cymbron et al., 2005). The Bantu expansion which started around 2000 BC was a major event in African history, and was probably responsible for the adoption of pastoralism (cattle, sheep and goats) by the Khoisan peoples of the Southern Africa region about 2 000 years ago (Hanotte et al., 2002) (Box 3). The origins of the indigenous pigs and chickens of the African continent remain largely undocumented. European colonization of the Americas led to the arrival of cattle, sheep, goats, pigs, horses and chickens in the New World. In the case of cattle there is genetic evidence for some African ancestry (Liron et al., 2006), which maybe a legacy of the slave trade between the two continents.

THE STATE OF AG R I CULT U R AL B I OD I VERSI T Y I N THE L I VESTOC K SEC TO R

Box 3 The history of African pastoralism Until recently, the history of African pastoralism was controversial and poorly understood. However, genetic marker analysis of indigenous cattle populations from all over the continent have now unravelled the major events in the history of pastoralism in Africa (Figure 4). The earliest African cattle originated within the continent, possibly as early as around 8000 BC. The exact centre(s) of domestication remain(s) unknown, but archaeological information suggests that it might have taken place in the northeastern part of the continent (Wendorf and Schild, 1994). These first African cattle were humpless Bos taurus animals. They initially dispersed north, as well as south to the borders of the tropical rainforests. Today, the only remaining descendants of these indigenous African taurine cattle are the trypanotolerant West African breeds (e.g. N’Dama and Baoulé), the Kuri, and the Sheko breed from Ethiopia. All these populations are now being intensively cross-bred with Zebu cattle (Bos indicus), and their unique genetic make-up is disappearing through unbalanced genetic admixture. Zebu cattle arrived in Africa much later. The earliest evidence for the presence of humped cattle is provided by Egyptian tomb paintings dating from the Twelfth Dynasty of the second millennium BC. It is probable that these animals were brought to Egypt in limited numbers as war treasure and, therefore are not connected to the later presence of Zebu cattle

in Africa. It is, however, thought that the Zebu was present in small numbers in the eastern part of the continent perhaps as early as 2 000 years ago as a result of early Arab contact or long-distance sea trade, and that this initial arrival resulted in the first introgression of Zebu genes into African taurine cattle. The major wave of Zebu arrival probably started with the Arab settlements along the East Coast of Africa from about the seventh century AD. The major inland dispersal of Zebu cattle probably followed the movement of pastoralists (e.g. Fulani throughout the Sahel), and was certainly accelerated by the rinderpest epidemics of the late nineteenth century. Southern Africa was the last area of the continent to acquire cattle pastoralism. Genetic data are now excluding a movement of cattle from the western part of the continent. It appears that herding spread southward from the Great Lakes region, which 2 000 years ago was the site of an Eastern Bantu core area. These farmers ultimately came into contact with San hunter–gatherers who acquired livestock from them. Influences from the Near East centre of cattle domestication are today found in the northeastern, northwestern and southern parts of the continent. The latter is probably a result of the settlement of European farmers in this part of the continent.

In Asia, the arrival of domestic livestock in the Japanese archipelago probably followed the establishment of farmers of Korean origin around 400 BC, but ancient influences from other geographic areas are also likely. In the Pacific, pigs and chickens had spread across western Polynesia by 900 to 700 BC, and the later Polynesian expansion carried these species as far as Rapa Nui (Easter Island) by 900 AD. Beside human migrations, ancient overland trading networks played an important role in the dispersion of livestock species. The domestication of livestock enabled large-scale overland trading

between civilizations, and livestock were themselves often a traded product. The main livestock species used as pack animals in the Old World were the donkey, horse, dromedary and Bactrian camel, and in South America, the llama. It is believed that domestication of the horse led to military expansion of horse-riding nomadic pastoralists in the Eurasian steppe, and subsequent dispersion of the species across the Old World. Bactrian camels were also used in warfare to a limited extent (Clutton-Brock, 1999), and the dromedary played an important role in the expansion of Arab civilization.

Adapted from Hanotte et al. (2002).

15

THE STATE OF THE WO RLD' S A N I MAL GENE T I C RESOU RC ES FO R FOOD A ND AG R I CULT U RE

PART 1

FIGURE 4 Origin and migration routes of domestic cattle in Africa

D

Centre of domestication Bos taurus (longhorn/shorthorn) 6000-2500 BC Bos taurus (African) 5000 BC-500 AD Bos indicus (Zebu - 1st wave) >2000 BC Bos indicus (Zebu - 2nd wave) >700 AD

Source: Graphics unit, ILRI (2006).

There is increasing evidence of the importance of ancient sea trading routes in the dispersion of livestock. For example, recent molecular genetic studies in cattle have revealed that Zebu animals were introduced into Africa via an Indian Ocean corridor rather than overland through the Isthmus of Suez or the Sinai Peninsula (Hanotte et al., 2002; Freeman et al., 2006). Similarly, both archaeological and genetic information suggest that the spread of pastoralism in the Mediterranean basin followed not only terrestrial costal routes, but also maritime routes (Zilhão, 2001; Beja-Pereira et al., 2006). A loss of diversity is to be expected following the dispersion and movement of livestock populations from their centres of origin. However, molecular

16

markers have revealed a more complex picture, with some movements resulting in an increase in diversity following admixture between populations originating from different centres of domestication. Additionally, detailed molecular studies indicate not only that cross-breeding between livestock populations was common, but also that genetic introgression from wild populations occurred after the initial domestication events. When they occurred outside the species’ geographic area of origin and after its initial dispersion, these wild introgressions may have resulted in localized livestock genetic populations with unique genetic backgrounds. Examples include local aurochs introgression in European (Götherström et

THE STATE OF AG R I CULT U R AL B I OD I VERSI T Y I N THE L I VESTOC K SEC TO R

al., 2005; Beja-Pereira et al., 2006) and possibly also in Asian cattle (Mannen et al., 2004). Unravelling the geographic pattern and history of the dispersal of livestock is essential to the identification of geographic areas with high levels of diversity, which are potential priority areas for conservation efforts. This requires extensive mapping of genetic diversity. Up to now, very few studies have been undertaken in this field. However, a recent study of cattle, covering Europe, Africa and West Asia, indicates that the highest degree of diversity is found in areas that are at the crossroads of admixture between populations from different centres of domestication (Freeman et al., 2006). An extensive survey of goat diversity in Europe and the Near and Middle East clearly indicates a geographical partitioning of goat diversity, with a large proportion of the genetic diversity among breeds explained by their geographic origins (Cañón et al., 2006). Today, local and regional, as well as transcontinental movement of livestock genotypes is accelerating as a result of the development and marketing of high-yielding breeds, new breeding technologies, and the increasing demand for livestock products. This modern dispersion, essentially restricted to a few breeds, and almost exclusively involving transfers from developed to developing countries, represents a major threat to the conservation and utilization of indigenous AnGR (see Section C for a further discussion of current gene flows).

5

Transformations in livestock following domestication

Mutation, selective breeding, and adaptation have shaped the diversity of livestock populations. The domestication process resulted in many changes some of which may still be ongoing. Particularly important have been morphological changes. Domestic animals are generally smaller than their wild ancestral counterparts (the notable exception being the chicken). Smaller animals are easier to manage and to handle, they may reach puberty sooner, and large flocks or herds can be kept more easily (Hall, 2004). The small West African cattle, sheep and dwarf goats are extreme examples of size reduction, possibly the result of genetic bottlenecks following adaptation to the tropical humid environment and its parasitic disease challenges. In some cases, human selection has deliberately resulted in extreme size differences – illustrated by the small size of the Shetland pony and the large size of the Shire horse (CluttonBrock, 1999). The body conformation of domestic animals may also be distinct from that of the wild ancestors – adapting, for example, to satisfy demand for meat products (e.g. European beef breeds), or to cope with new environmental pressures (e.g. Sahelian goats). Selection for muscular mass has often resulted in greater muscular development of the hind quarters relative to the shoulders (Hall, 2004). An extreme example of selection for muscular mass is the double-muscling trait observed in some European beef breeds, and in some sheep and pigs breeds. In cattle, the trait results from mutation at a single gene – the myostatin gene (Grobet et al., 1998). In sheep, it involves the callipyge gene (Cockett et al., 2005). The pattern of fat deposition may also show changes following domestication. For example, reduced predation has encouraged fat deposition in domestic poultry. In domesticated mammals, the hump of the Zebu and the tails of fat-tailed and fat-rumped sheep are striking examples of selection for fat deposition. This exaggerated fat deposition may be quite ancient, with fat-

17

THE STATE OF THE WO RLD' S A N I MAL GENE T I C RESOU RC ES FO R FOOD A ND AG R I CULT U RE

PART 1

tailed sheep already common in western Asia by 3000 BC, and humped cattle depicted on cylinder seals from the ancient civilizations of MohenjoDaro and Harappa in the Indus Valley about 2500 to 1500 BC (Clutton-Brock 1999). Great variation is found in the wool and hair coats of most domestic species. For example, sheep breeds of alpine regions have particularly thick woolly coats, while breeds from the African Sahel lack wool. It is probable that these changes were the result of mutations followed by artificial selection, perhaps as early as 6000 BC, as illustrated by a statuette of a woolly sheep found in the Islamic Republic of Iran (Clutton-Brock, 1999). Coat and plumage coloration were also selected by the environment, with light coloured animals being more adapted to hotter environments and dark coloured animals to cooler environments (Hall 2004). Coat colours have also been influenced by cultural selection. Livestock breeders in the developed world often favour uniformity in coat colour, but in the tropics, diversity in coat colour may be preferred for ceremonial reasons, or simply to facilitate the identification of individual animals. An illustration of the latter is the great diversity in coat colours and patterns observed among the Nguni cattle of the Zulu people (Poland et al., 2003). It is important to realize that local adaptation, human and/or natural selection will not always result in reduced genetic variation or functional diversity in the livestock population. For example, natural selection may favour adaptive diversity within herds kept in changing environments (e.g. as a result of climatic variation). A recent study of the genetic diversity of the six most important milk proteins in cattle revealed higher diversity in a relatively restricted geographic area of northern Europe, with selection pressure imposed by early (milk drinking) pastoralists being the most likely explanation (Beja-Pereira et al, 2003).

18

6

Conclusions

Understanding of the origin and subsequent history and evolution of AnGR diversity is essential to the design of sustainable conservation and utilization strategies. Livestock diversity originates from the wild ancestors, and was subsequently shaped through the processes of mutation, genetic drift, and natural and human selection. Only a subset of the diversity present in the ancestral species survived in the domestic counterparts. However, domestic livestock diversity has been continuously evolving. Reshuffling of genes at each generation, mutation, and cross-breeding or admixture of different gene pools has offered new opportunities for natural and human selection. This has been the basis of the enormous gains in output achieved in commercial breeds, and of the adaptation of indigenous livestock to highly diverse and challenging environments. However, the world’s livestock diversity is currently shrinking – with rapid and uncontrolled loss of unique and often uncharacterized AnGR. If a breed or population becomes extinct, this means the loss of its unique adaptive attributes, which are often under the control of many interacting genes, and are the results of complex interactions between the genotype and the environment.

THE STATE OF AG R I CULT U R AL B I OD I VERSI T Y I N THE L I VESTOC K SEC TO R

References Beja-Pereira, A., Caramelli, D., Lalueza-Fox, C., Vernesi, C., Ferrand, N., Casoli, A., Goyache, F., Royo, L.J., Conti, S., Lari, M., Martini, A., Ouragh, L., Magid, A., Atash, A., Zsolnai, A., Boscato, P., Triantaphylidis, C., Ploumi, K., Sineo, L., Mallegni, F., Taberlet, P., Erhardt, G., Sampietro, L., Bertranpetit, J., Barbujani, G., Luikart, G. & Bertorelle, G. 2006. The origin of European cattle: evidence from modern and ancient DNA. Proceedings of the National Academy of Sciences USA, 103(21): 8113–8118. Beja-Pereira, A., England, P.R., Ferrand, N., Jordan, S., Bakhiet, A.O., Abdalla, M.A., Maskour, M., Jordana, J., Taberlet, P. & Luikart, G. 2004. African origin of the domestic donkey. Science, 304(5678): 1781. Beja-Pereira, A., Luikart, G., England, P.R., Bradley, D.G., Jann, O.C., Bertorelle, G., Chamberlain, A.T., Nunes, T.P., Metodiev, S., Ferrand, N. & Erhardt, G. 2003. Gene-culture coevolution between cattle milk protein genes and human lactase genes. Nature Genetics, 35(4): 311–313. Bogucki, P. 1996. The spread of early farming in Europe. American Science, 84: 242–253. BOSTID. 1991. Microlivestock: little-known small animals with a promising economic future. Washington DC. National Academic Press. Bradley, D.G., MacHugh, D.E., Cunningham, P. & Loftus, R.T. 1996. Mitochondrial DNA diversity and the origins of African and European cattle. Proceedings of the National Academy of Sciences USA, 93(10): 5131–5135. Bradley, D.G. & Magee, D. 2006. Genetics and the origins of domestic cattle. In M.A. Zeder, E. Emshwiller, B.D. Smith & D.G. Bradley, eds. Documenting domestication: new genetics and archaeological paradigm, pp. 317–328. California, USA. University of California Press. Bruford, M.W., Bradley, D.G. & Luikart, G. 2003. DNA markers reveal the complexity of livestock domestication. Nature Reviews Genetics, 4(11): 900–909.

Bulliet, R.W. 1975. The Camel and the wheel. Massachusetts, USA. Harvard University Press. Cañón, J., Garcia, D., Garcia-Atance, M.A., ObexerRuff, G., Lenstra, J. A., Ajmone-Marsan, P., Dunner, S. & the ECONOGENE Consortium. 2006. Geographical partitioning of goat diversity in Europe and the Middle East. Animal Genetics, 37(4), 327–334. Clutton-Brock, J. 1999. A natural history of domesticated mammals. 2nd Edition. Cambridge, UK. Cambridge University Press. Cockett, N.E., Smit, M.A., Bidwell, C.A., Segers, K., Hadfield, T.L., Snowder, G.D., Georges, M. & Charlier, C. 2005. The callipyge mutation and other genes that affect muscle hypertrophy in sheep. Genetic Selection and Evolution, 37(Suppl 1): 65–81. Compagnoni, B. & Tosi, M. 1978. The camel: its distribution and state of domestication in the Middle East during the third millennium B.C. in light of finds from Shahr-i Sokhta. In R.H. Meadow, & M.A Zeder, eds. Approaches to faunal analysis in the Middle East. Peabody Museum Bulletin 2, pp. 91–103. Cambridge MA, USA. Peabody Museum. Cymbron, T., Freeman, A.R., Malheiro, M.I, Vigne, J.-D. & Bradley, D.G. 2005. Microsatellite diversity suggests different histories for Mediterranean and Northern European cattle populations. Proceedings of the Royal Society of London B, 272: 1837–1843. Diamond, J. 1999. Guns, germs and steel: the fates of human societies. New York, USA. Norton. Diamond, J. 2002. Evolution, consequences and future of plant and animal domestication. Nature, 418: 700–707. Diamond, J. & Bellwood, P. 2003. Farmers and their languages: the first expansions. Science, 300: 597–603. Dobney, K. & Larson, G. 2006. Genetics and animal domestication: new windows on an elusive process. Journal of Zoology, 269: 261–271.

19

THE STATE OF THE WO RLD' S A N I MAL GENE T I C RESOU RC ES FO R FOOD A ND AG R I CULT U RE

PART 1

FAO. 2005. Genetic characterization of livestock populations and its use in conservation decision making, by O. Hannotte & H. Jianlin. In J. Ruane & A. Sonnino, eds. The role of biotechnology in exploring and protecting agricultural genetic resources, pp. 89–96. Rome. (also available at www.fao.org/docrep/009/ a0399e/a0399e00.htm) Felius, M. 1995. Cattle breeds – an encyclopedia. Doetinchem, the Netherlands. Misset. Fernández, H., Hughes, S., Vigne, J.-D., Helmer, D., Hodgins, G., Miquel, C., Hänni, C., Luikart, G. & Taberlet, P. 2006. Divergent mtDNA lineages of goats in an early Neolithic site, far from the initial domestication areas. Proceedings of the National Academy of Sciences USA, 103(42): 15375–15379. Freeman, A.R., Bradley, D.G., Nagda, S., Gibson, J.P. & Hanotte, O. 2006. Combination of multiple microsatellite datasets to investigate genetic diversity and admixture of domestic cattle. Animal Genetics, 37(1): 1–9. Fumihito, A., Miyake, T., Sumi, S., Takada, M., Ohno, S. & Kondo, N. 1994. One subspecies of the red junglefowl (Gallus gallus gallus) suffices as the matriarchic ancestor of all domestic breeds. Proceedings of the National Academy of Sciences USA, 91(26): 12505–12509. Fumihito, A., Miyake, T., Takada, M., Shingu, R., Endo, T., Gojobori, T., Kondo, N. & Ohno, S. 1996. Monophyletic origin and unique dispersal patterns of domestic fowls. Proceedings of the National Academy of Sciences USA, 93(13): 6792–6795. Götherström, A., Anderung, C., Hellborg, C., Elburg, R., Smith, C., Bradley, D.G. & Ellegren, H. 2005. Cattle hybridization in the Near East was followed by hybridization with auroch bulls in Europe. Proceedings of the Royal Society of London B, 272: 2345–2350. Grobet, L., Poncelet, D., Royo, L.J., Brouwers, B., Pirottin, D., Michaux, C., Menissier, F., Zanotti, M., Dunner, S. & Georges, M. 1998. Molecular definition of an allelic series of mutations disrupting the myostatin function and causing double-muscling in cattle. Mammalian Genome, 9(3): 210–213.

20

Guiffra, E., Kijas, J.M.H., Amarger, V., Calborg, Ö., Jeon, J.T. & Andersson, L. 2000. The origin of the domestic pigs : independent domestication and subsequent introgression. Genetics, 154(4): 1785–1791. Guo, S., Savolainen, P., Su, J., Zhang, Q., Qi, D., Zhou, J., Zhong, Y., Zhao, X. & Liu, J. 2006. Origin of mitochondrial DNA diversity in domestic yak. BMC Evolutionary Biology, 6: 73. Hall, S.J.G. 2004. Livestock biodiversity: genetic resources for the farming of the future. Oxford, UK. Blackwell Science Ltd. Hanotte, O., Bradley, D.G., Ochieng, J., Verjee, Y., Hill, E.W. & Rege, J.E.O. 2002. African pastoralism: genetic imprints of origins and migrations. Science, 296(5566): 336–339. Hanotte, O. & Mensah, G.A. 2002. Biodiversity and domestication of ‘non-conventional’ species: a worldwide perspective. Seventh World Congress on Genetics Applied to Livestock Production, 19–23 August 2002, Montpellier, France. 30: 543–546. Hanotte, O., Toll J., Iniguez L. & Rege, J.E.O. 2006. Farm animal genetic resources: why and what do we need to conserve. Proceeding of the IPGRI–ILRI– FAO–CIRAD workshop: Option for in situ and ex situ conservation of AnGR, 8–11 November 2005, Montpellier, France. Hiendleder, S., Mainz, K., Plante, Y. & Lewalski, H. 1998. Analysis of mitochondrial DNA indicates that the domestic sheep are derived from two different ancestral maternal sources: no evidences for the contribution from urial and argali sheep. Journal of Heredity, 89: 113–120. Higham, C. 1975. Non Nok Tha, the funeral remains from the 1966 and 1968 excavations at Non Nok Tha Northeastern Thailand. Studies in Prehistoric Anthropology Volume 6. Otago, New Zealand. University of Otago. Jansen, T., Foster, P., Levine, M.A., Oelke, H., Hurles, M., Renfrew, C., Weber, J. & Olek, K. 2002. Mitochondrial DNA and the origins of the domestic horse. Proceedings of the National Academy of Science USA, 99(16): 10905–10910.

THE STATE OF AG R I CULT U R AL B I OD I VERSI T Y I N THE L I VESTOC K SEC TO R

Jianlin H., Quau J., Men Z., Zhang Y. & Wang W. 1999. Three unique restriction fragment length polymorphisms of EcoR I, Pvu II and Sca I digested mitochondrial DNA of wild Bactrian camel (Camelus bactrianus ferus) in China. Journal of Animal Science, 77: 2315–2316. Joshi, M.B., Rout, P.K., Mandal, A.K., Tyler-Smith, C., Singh, L. & Thangaray, K. 2004. Phylogeography and origins of Indian domestic goats. Molecular Biology and Evolution, 21(3): 454–462. Kadwell, M., Fernández, M., Stanley, H.F., Baldi, R., Wheeler, J.C., Rosadio, R. & Bruford, M.W. 2001. Genetic analysis reveals the wild ancestors of the llama and alpaca. Proceedings of the Royal Society of London B, 268: 2675–2584. Larson, G., Dobney, K., Albarella, U., Fang, M., Matisoo-Smith, E., Robins, J., Lowden, S., Finlayson, H., Brand, T., Willerslev, E., RowleyConwy, P., Andersson, L. & Cooper, A. 2005. Worldwide phylogeography of wild boar reveals multiple centers of pig domestication. Science, 307(5715): 1618–1621. Liron, J.P., Bravi, C.M., Mirol, P.M., Peral-Garcia, P. & Giovambattista, G. 2006. African matrilineages in American Creole cattle: evidence of two independent continental sources. Animals Genetics, 37(4): 379–382. Liu, Y.P., Wu, G.-S., Yao, Y.G., Miao, Y.W., Luikart, G., Baig, M., Beja-Pereira, A., Ding, Z.L., Palanichamy, M.G. & Zhang, Y.-P. 2006. Multiple maternal origins of chickens: out of the Asian jungles. Molecular Phylogenetics and Evolution, 38(1): 12–19. Loftus, R.T., MacHugh, D.E., Bradley, D.G., Sharp, P.M. & Cunningham, P. 1994. Evidence for two independent domestication of cattle. Proceedings of the National Academy of Sciences USA, 91(7): 2757–2761. Luikart, G.L., Gielly, L., Excoffier, L., Vigne, J-D., Bouvet, J. & Taberlet, P. 2001. Multiple maternal origins and weak phylogeographic structure in domestic goats. Proceedings of the National Academy of Sciences USA, 98(10): 5927–5930.

Mannen, H., Kohno, M., Nagata, Y., Tsuji, S., Bradley, D.G., Yeao, J.S., Nyamsamba, D., Zagdsuren, Y., Yokohama, M., Nomura, K. & Amano, T. 2004. Independent mitochondrial DNA origin and historetical genetic differentiation in North Eastern Asian cattle. Molecular Phylogenetic and Evolution, 32(2): 539–544. Mignon-Grasteau, S., Boissy, A., Bouix, J., Faure, J.-M., Fisher, A.D., Hinch, G.N., Jensen, P., Le Neindre, P., Mormède, P., Prunet, P., Vandeputte, M. & Beaumont, C. 2005. Genetics of adaptation and domestication in livestock. Livestock Production Science, 93(1): 3–14. Olsen, S.L. 2006. Early horse domestication on the Eurasian steppe. In M.A. Zeder, E. Emshwiller, B.D. Smith & D.G. Bradley, eds. Documenting domestication: new genetics and archaeological paradigms, pp. 245–269. California, USA. University of California Press. Pedrosa, S., Uzun, M., Arranz, J.J., Guttiérrez-Gil, B., San Primitivo, F. & Bayon, Y. 2005. Evidence of three maternal lineages in Near Eastern sheep supporting multiple domestication events. Proceedings of the Royal Society of London B, 272(1577): 2211–2217. Peters, J., Helmer, D., von den Driesch, A. & Segui, S. 1999. Animal husbandry in the northern Levant. Paléorient, 25: 27–48. Peters, J. & von den Driesch, A. 1997. The twohumped camel (Camelus bactrianus): new light on its distribution management and medical treatment in the in the past. Journal of Zoology, 242: 651–679. Poland, M., Hammond-Tooke, D. & Leigh, V. 2003. The abundant herds: a celebration of the cattle of the Zulu people. Vlaeberg, South Africa. Fernwood Press. Qi, X. 2004. Genetic diversity, differentiation and relationship of domestic yak populations: a microsatellite and mitochondrial DNA study. Lanzhou University, China. (PhD Thesis) Ryder, M.L. 1984. Sheep. In I.L. Mason, ed. Evolution of domesticated animals, pp. 63–65. London. Longman.

21

THE STATE OF THE WO RLD' S A N I MAL GENE T I C RESOU RC ES FO R FOOD A ND AG R I CULT U RE

PART 1

Skjenneberg, S. 1984. Reindeer. In I.L. Mason, ed. Evolution of domesticated animals, pp. 128–138. London. Longman. Sultana, S., Mannen, H. & Tsuji, S. 2003. Mitochondrial DNA diversity of Pakistani goats. Animal Genetics, 34(6): 417–421. Tanaka, K., Solis, C.D., Masangkay, J.S., Maeda, K., Kawamoto, Y. & Namikawa, T. 1996. Phylogenetic relation among all living species of the genus Bubalus based on DNA sequences of the cytochrome B gene. Biochemical Genetics, 34(11–12): 443–452. Tapio, M., Marzanov, N., Ozerov, M., C’inkulov, M., Gonzarenko, G., Kiselyova, T., Murawski, M., Viinalass, H. & Kantanen, J. 2006. Sheep mitochondrial DNA in European Caucasian and Central Asian areas. Molecular Biology and Evolution, 23(9): 1776–1783. Vilà, C., Leonard, J.A., Götherström, S., Marklund, S., Sanberg, K., Lindén, K., Wayne, R.K. & Ellegren, H. 2001. Widespread origins of domestic horse lineages. Science, 291(5503): 474–477. Vilà, C., Leonard, J.A. & Beja-Pereira, A. 2006. Genetic documentation of horse and donkey domestication. In M.A. Zeder, E. Emshwiller, B.D. Smith & D.G. Bradley, eds. Documenting domestication: new genetics and archaeological paradigms, pp. 342–353. California, USA. University of California Press. Wayne, R.K., Leonard, J.A. & Vilà, C. 2006. Genetic analysis of dog domestication. In M.A. Zeder, E. Emshwiller, B.D. Smith & D.G. Bradley, eds. Documenting domestication: new genetics and archaeological paradigms, pp. 279–293. California, USA. University of California. Wendorf, F. & Schild, R. 1994. Are the early Holecene cattle in the Eastern Sahara domestic or wild? Evolutionary Anthropology, 3: 118–128. West, B. & Zhou, B-X. 1988. Did chickens go north? New evidence for domestication. Journal of Archaeological Science, 15: 515–533.

22

Wheeler, J.C., Chikni, L. & Bruford, M.W. 2006. Genetic analysis of the origins of domestic South American Camelids. In M.A. Zeder, E. Emshwiller, B.D. Smith & D.G. Bradley, eds. Documenting domestication: new genetics and archaeological paradigms, pp. 279–293. California, USA. University of California Press. Zeder, M.A., Emshwiller, E., Smith, B.D. & Bradley, D.G. 2006. Documenting domestication: the intersection of genetics and archaeology. Trends in Genetics, 22(3): 139–155. Zeder, M.A. & Hesse, B. 2000. The initial domestication of goats (Capra hircus) in the Zagros mountains 10,000 years ago. Science, 287(5461): 2254–2257. Zeuner, F.E. 1963. A history of domesticated animals. London. Hutchinson. Zilhão, J. 2001. Radiocarbon evidences for maritime pioneer colonization at the orign of farming in West Mediterranean Europe. Proceedings of the National Academy of Sciences USA, 98(24): 14180–14185.

THE STATE OF AG R I CULT U R AL B I OD I VERSI T Y I N THE L I VESTOC K SEC TO R

Section B

Status of animal genetic resources 1

Introduction

This section presents a global overview of the diversity and status of AnGR. The analysis is based on FAO’s Global Databank for Animal Genetic Resources for Food and Agriculture (Global Databank), as it is the only such resource that provides worldwide coverage. It serves as an updated (but condensed) version of the World Watch List for Domestic Animal Diversity2 (WWL– DAD), the previous (third) edition of which was published in 2000. Box 4 outlines changes in the approach to reporting and data analysis that have been introduced for the State of the World’s Animal Genetic Resources for Food and Agriculture (SoWAnGR) preparation process. The section begins by describing the state of reporting on AnGR, and the progress made during the period December 1999 to January 2006. A description of the current 2

FAO/UNEP 2000. World watch list for domestic animal diversity, 3rd edition, edited by B.D. Scherf, Rome. (also available at http://www.fao.org/dad-is).

regional distribution of livestock species and breeds is then presented, followed by an overview of the risk status of the world’s livestock breeds. Finally, trends in risk status over this six year period are assessed.

2

State of reporting

The total number of breed records in the Global Databank has increased greatly since the publication of the WWL–DAD:3 (Table 5). The total number of entries rose from 6 379 in December 1999 to 14 017 in January 2006. The increase was particularly marked in the case of avian breed populations, for which the number of records increased from 1 049 to 3 505. In the case of mammalian species the number rose from 5 330 to 10 512. Nearly all breed populations reported (94 percent) are domesticated livestock, only 1 percent are feral, and less than 1 percent

TABLE 5 Status of information recorded in the Global Databank for Animal Genetic Resources Year of analysis

Mammalian species Number of national breed populations

Avian species

% with population data

Number of national breed populations

1993

2 719

53

-

1995

3 019

73

1999

5 330

2006

10 512

% with population data

Countries covered

-

131

863

85

172

63

1 049

77

172

43

3 505

39

182*

*No data recorded for Andorra, Brunei Darussalam, Gaza Strip, Holy See, Liechtenstein, Marshall Islands, Federated States of Micronesia, Monaco, Nauru, Qatar, San Marino, Singapore, Timor-Leste, United Arab Emirates, West Bank, Western Sahara.

23

THE STATE OF THE WO RLD' S A N I MAL GENE T I C RESOU RC ES FO R FOOD A ND AG R I CULT U RE

PART 1

are wild populations (for the remaining 4 percent no specification was given). While the number of breeds recorded has increased, the percentage of breeds for which population data are available, decreased from 77 to 39 percent for avian breeds, and from 63 to 43 percent for mammalian breeds (Table 5 and Figure 5). Furthermore, where population figures are reported, they may not have been updated recently. The large discrepancy between

the number of breed entries and the number for which population data are available is in part accounted for by the fact that much of the latest data entered into the Global Databank were extracted from Country Reports. These reports often mention the existence of breeds, but do not include details of population size. Before analysis of the global state of breed diversity and risk status could be undertaken, some adjustments to the raw figures for the

Box 4 What is new compared to the World Watch List for Domestic Animal Diversity? In 1991, FAO initiated global breed surveys to report on the seven major mammalian domestic animal species (ass, buffalo, cattle, goat, horse, pig and sheep). Additional surveys were initiated in 1993 to include yaks, the six camelid species and the 14 major avian species. Collection of data for deer species and rabbits followed, and these species were included in the third edition of the World Watch List for Domestic Animal Diversity (WWL–DAD:3) published in 2000. In order to produce a more complete inventory, FAO provided, during 2005, for the extraction of breed-related data from 169 Country Reports, and the entry of these data into the Global Databank for Animal Genetic Resources. Subsequently, National Coordinators (NCs) were requested to validate and further complete their national breed databanks. The WWL–DAD:3 (2000) was criticized for overestimating the number of breeds categorized as being “at risk”. This overestimation occurred because risk status was assigned to each national breed population based on the population size in the particular country. Thus, in the case of breeds that occur in more than one country, there was a danger that the categorization was not a true reflection of risk status. This problem had previously been recognized, but at the time the emphasis of reporting was on local breeds. For the SoW-AnGR process, countries decided to consider all their AnGR (both local and imported). The number of breeds wrongly categorized as being at risk would, therefore, have

24

greatly increased. The new analysis attempts to correct this bias by linking national breed populations that belong to a common gene pool. This linkage was implemented based on expert knowledge and revised by NCs. However, a clear definition of what constitutes a common gene pool is still lacking. The linked breeds are referred to as transboundary breeds (Box 5). Risk status for these breeds is estimated based on the overall number of animals belonging to the breed in question. The method of assessing breed diversity at regional and global levels has also been adapted: at the regional level, breeds that reside in more than one country, but only within the SoW-AnGR region in question, are now counted only once for the region regardless of how many national-level populations there may be. International transboundary breeds, which occur in many regions, are counted only once at the global level. When comparing the WWL–DAD:3 with the figures provided in this Report, it must be noted that the classification of regions has also been changed. Southwest Pacific and Asia are here considered to be separate regions, while “Asia and the Pacific” was considered a single region in WWL–DAD 3. Moreover, it should be noted that the regional classification used in this Report is also different from the standard FAO regional classification.

THE STATE OF AG R I CULT U R AL B I OD I VERSI T Y I N THE L I VESTOC K SEC TO R

Box 5 Glossary: populations, breeds, regions Wild populations: represent either wild relatives of domesticated livestock, wild populations that are used for food and agriculture, or populations undergoing domestication.

Transboundary breeds: breeds that occur in more than one country. These are further differentiated as: – Regional transboundary breeds: transboundary breeds that occur only in one of the seven SoW-AnGR regions. – International transboundary breeds: transboundary breeds that occur in more than one region.

Feral populations: animals are considered to be feral if they or their ancestors were formerly domesticated, but they are now living independently of humans; for example, dromedaries in Australia.

SoW-AnGR regions: seven regions were defined for the SoW-AnGR: Africa, Asia, Europe and the Caucasus, Latin America and the Caribbean, the Near and Middle East, North America, and the Southwest Pacific.

Local breeds: breeds that occur only in one country.

FIGURE 5 Proportion of national breed populations for which population figures have been reported

Southwest Pacific

112

20

Avian

310

88

Mammalian

89

Avian

2

North America

Near & Middle East Latin America & the Caribbean

Europe & the Caucasus

126

200

Mammalian

46

14

Avian Mammalian

127

172 309

49

Avian Mammalian

1 428

212

Avian

617

865

Mammalian

1 144

2 344

Avian

352

248

Asia Mammalian Avian

851

1 080 276

75

Africa Mammalian 0%

1 077

519 20%

40%

with population data

60%

80%

100%

without population data

25

THE STATE OF THE WO RLD' S A N I MAL GENE T I C RESOU RC ES FO R FOOD A ND AG R I CULT U RE

PART 1

number of breed populations were required. Four hundred and eighty entries classified as “strains” or “lines” were excluded from the analysis (in the case of avian species, further validation by national and regional experts to link lines and strains to the respective breeds is needed). Furthermore, 209 breed populations that obviously belonged to the same breed, but had been reported twice from the same country were excluded. These adjustments left a total of 13 328 breed populations for inclusion in the analysis of diversity and risk status.

Slightly more than half of the total number of recorded national breed populations (6 792 entries) occur in more than one country. These breed populations have been linked and are defined as “transboundary” breeds (Box 5). The risk status assigned to a transboundary breed takes into account all reported populations for the breed in question. Breed populations occurring only in one country are defined as “local” breeds. Transboundary breeds are classified as either “regional” or “international”, depending on the extent of their distribution (Box 5).

TABLE 6 Distribution of mammalian species by region Mammalian species

Africa

Asia

Europe & the Caucasus

Latin America & the Caribbean

Near & Middle East

North America

Southwest Pacific

percentage of countries in a region reporting breed-related information for the species Buffalo

8

57

25

27

25

0

8

Cattle

98

96

100

94

75

100

77

0

32

2

0

0

0

0 69

Yak Goat

96

96

93

94

83

100

Sheep

92

86

100

91

100

100

31

Pig

70

82

91

91

8

100

92

Ass

38

46

36

39

50

50

8

Horse

46

93

91

64

58

100

23

0

25

5

0

0

0

0

32

25

2

0

58

0

8

2

0

0

12

0

0

8

Bactrian camel Dromedary Alpaca Llama

0

0

0

15

0

0

0

Guanaco

0

0

0

9

0

0

0

Vicuña

0

0

0

12

0

0

0

Deer*

2

25

14

9

0

50

15

Rabbit

38

39

39

48

8

0

0

Guinea pig

8

0

0

15

0

0

0

Dog

2

7

5

0

0

0

0

Shading: purple: ≥50% of countries; green: 0 to 35 35 to 250 > 250 Data not available

Source: FAOSTAT.

85

THE STATE OF THE WO RLD' S A N I MAL GENE T I C RESOU RC ES FO R FOOD A ND AG R I CULT U RE

PART 1

FIGURE 35 Net exports – eggs

Eggs (poultry) Net exports in 2003 [1 000 US$] < -4 000 -4 000 to -1 200 -1 200 to -400 -400 to -80 -80 to 0 > 0 to 800 800 to 7 000 > 7 000 Data not available

Source: FAOSTAT.

export commodities in many countries. Trade in livestock products is growing, but faces a number of constraints – particularly associated with animal health. The countries of the world can be distinguished according to whether they are net exporters or net importers of particular animal products. Figures 33, 34 and 35 show the export/ import status of countries for meat, milk and eggs respectively. Brazil and the southern countries of South America are net exporters of meat, as are the countries of North America; Australia and New Zealand; a number African countries (most notably Botswana and Namibia); China, India and several other Asian countries; as well as many European countries. In the case of milk, long-standing net exporters such as Argentina, Australia and New Zealand, have been joined in recent years by new exporting countries such as Colombia, India and Kyrgyzstan. Net exporters of eggs can be found in all regions of the world. In Asia, for example, major net exporters include China, India, the Islamic Republic of Iran and Malaysia. The largest

86

net exporter of eggs in the Africa region is South Africa, but there are a number of other such countries including Ethiopia, Zambia and Zimbabwe. In Latin America and the Caribbean, Colombia and Peru have in recent years become net exporters of eggs, as has Egypt in the Near and Middle East.

5

Production of fibre, skins, hides and pelts

Livestock fibres, hides, skins and pelts are also important products. Although the world’s sheep industry has over recent years seen a shift in orientation away from wool production and towards meat, wool remains an important product in many countries. The Southwest Pacific is the region of the world that produces the most wool (Table 28). China, the Islamic Republic of Iran, the United Kingdom and other countries with large sheep populations are also major producers of wool, but it is often of secondary importance to

THE STATE OF AG R I CULT U R AL B I OD I VERSI T Y I N THE L I VESTOC K SEC TO R

TABLE 28 Production of fibres, skins and hides (1000 tonnes/year) Asia

Cattle Hides, Fresh

515.5

2 576.7

1 377.8

1 809.0

119.7

1 157.7

304.1

Goatskins, Fresh

112.2

727.9

30.6

23.2

64.9

0.01

5.4

Sheepskins, Fresh

0.05

0.03

0.06

0.03

0.01

10 000) that can be separated and visualized in a single experiment. Protein spots are cut from the gel, followed by proteolytic digestion, and proteins are then identified using mass spectrometry (Aebersold and Mann, 2003). However, standardization and automation of two-dimensional gel electrophoresis has proved difficult, and the use of the resulting protein patterns as proteomic reference maps has only been successful in a few cases. A complementary technique, liquid chromatography, is easier to automate, and it can be directly coupled to mass spectrometry. Affinity-based proteomic methods that are based on microarrays are an alternative approach to protein profiling (Lueking et al., 2003), and can also be used to detect protein– protein interactions. Such information is essential for algorithmic modelling of biological pathways. However, binding specificity remains a problem in the application of protein microarrays, because cross-reactivity cannot accurately be predicted. Alternative approaches exist for detecting protein–protein interactions such as the two hybrid system (Fields and Song, 1989). However, none of the currently used methods allow the quantitative detection of binding proteins, and it remains unclear to what extent the observed interactions are likely to represent the physiological protein–protein interactions. Array-based methods have also been developed for detecting DNA–protein interaction in vitro and in vivo (see Sauer et al., 2005, for a review), and identifying unknown proteins binding to gene regulatory sequences. DNA microarrays are employed effectively for screening nuclear extracts for DNA-binding complexes, whereas

371

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

protein microarrays are mainly used for identifying unknown DNA-binding proteins at proteomewide level. In the future, these two techniques will reveal detailed insights into transcriptional regulatory networks. Many methods of predicting the function of a protein are based on its homology to other proteins and its location inside the cell. Predictions of protein functions are rather complicated, and also require techniques to detect protein–protein interactions, and to detect the binding of proteins to other molecules, because proteins fulfil their functions in these binding processes.

4

The role of bioinformatics

Developing high-throughput technologies would be useless without the capacity to analyse the exponentially growing amount of biological data. These need to be stored in electronic databases (Box 78) associated with specific software designed to permit data update, interrogation and retrieval. Information must be easily accessible and interrogation-flexible, to allow the retrieval of information, that can be analysed to unravel metabolic pathways and the role of the proteins and genes involved. Bioinformatics is crucial to combine information from different sources and generate new knowledge from existing data. It also has the potential to simulate the structure, function and dynamics of molecular systems, and is therefore helpful in formulating hypotheses and driving experimental work.

5

Conclusions

Molecular characterization can play a role in uncovering the history, and estimating the diversity, distinctiveness and population structure of AnGR. It can also serve as an aid in the genetic management of small populations, to avoid excessive inbreeding. A number of investigations

372

Box 78 Databases of biological molecules A number of databases exist which collect information on biological molecules: DNA sequence databases: • European Molecular Biology Lab (EMBL): http:// www.ebi.ac.uk/embl/index.html • GenBank: http://www.ncbi.nlm.nih.gov/ • DNA Data Bank of Japan (DDBJ): http://www. ddbj.nig.ac.jp Protein databases: • SWISS-PROT: http://www.expasy.ch/sprot/sprottop.html • Protein Information Resource (PIR): http://pir. georgetown.edu/pirwww/ • Protein Data Bank (PDB): http://www.rcsb.org/ pdb/ Gene identification utility sites Bio-Portal • GenomeWeb: http://www.hgmp.mrc.ac.uk/ GenomeWeb/nuc-geneid.html • BCM Search Launcher: http://searchlauncher. bcm.tmc.edu/ • MOLBIOL: http://www.molbiol.net/ • Pedro’s BioMolecular Research tools: http:// www.biophys.uni-duesseldorf.de/BioNet/Pedro/ research_tools.html • ExPASy Molecular Biology Server: http://www. expasy.ch/ Databases of particular interest for domestic animals: http://locus.jouy.inra.fr/cgi-bin/bovmap/intro.pl http://www.cgd.csiro.au/cgd.html http://www.ri.bbsrc.ac.uk/cgi-bin/arkdb/browsers/ http://www.marc.usda.gov/genome/genome.html http://www.ncbi.nlm.nih.gov/genome/guide/pig/ http://www.ensembl.org/index.html http://www.tigr.org/ http://omia.angis.org.au/ http://www.livestockgenomics.csiro.au/ibiss/ http://www.thearkdb.org/ http://www.hgsc.bcm.tmc.edu/projects/bovine/

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

have described within and between-population diversity – some at quite a large scale. However, these studies are fragmented and difficult to compare and integrate. Moreover, a comprehensive worldwide survey of relevant species has not been carried out. As such, it is of strategic importance to develop methods for combining existing, partially overlapping datasets, and to ensure the provision of standard samples and markers for future use as worldwide references. A network of facilities collecting samples of autochthonous germplasm, to be made available to the scientific community under appropriate regulation, would facilitate the implementation of a global survey. Marker technologies are evolving, and it is likely that microsatellites will increasingly be complemented by SNPs. These markers hold great promise because of their large numbers in the genome, and their suitability for automation in production and scoring. However, the efficiency of SNPs for the investigation of diversity in animal species remains to be thoroughly explored. The subject should be approached with sufficient critical detachment to avoid the production of biased results. Methods of data analysis are also evolving. New methods allow the study of diversity without a priori assumptions regarding the structure of the populations under investigation; the exploration of diversity to identify adaptive genes (e.g. using population genomics, see Box 77); and the integration of information from different sources, including socio-economic and environmental parameters, for setting conservation priorities (see Section F). The adoption of a correct sampling strategy and the systematic collection of phenotypic and environmental data, remain key requirements for exploiting the full potential of new technologies and approaches. In addition to neutral variation, research is actively seeking genes that influence key traits. Disease resistance, production efficiency, and product quality are among the traits having high priority. A number of strategies and new

high-throughput –omics technologies are used to this end. The identification of QTN offers new opportunities and challenges for AnGR management. Information on adaptive diversity complements that on phenotypic and neutral genetic diversity, and can be integrated into AnGR management and conservation decision-tools. The identification of unique alleles or combinations of alleles for adaptive traits in specific populations may reinforce the justification for their conservation and targeted utilization. Gene assisted selection also has the potential to decrease the selection efficiency gap currently existing between large populations raised in industrial production systems, and small local populations, where population genetic evaluation systems and breeding schemes cannot be effectively applied. Marker and gene assisted selection may not, however, always represent the best solution. These options need to be evaluated and optimized on a case-by-case basis, taking into account short and long-term effects on population structure and rates of inbreeding, and cost and benefits in environmental and socio-economic terms – in particular impacts on people’s livelihoods. As in the case of other advanced technologies, it is highly desirable that benefits of scientific advances in the field of molecular characterization are shared across the globe, thereby contributing to an improved understanding, utilization and conservation of the world’s AnGR for the good of present and future human generations.

373

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

Box 79 Glossary: molecular markers For the purpose of this section the following definitions are used: Candidate gene: any gene that could plausibly cause differences in the observable characteristics of an animal (e.g. in disease resistance, milk protein production or growth). The gene may be a candidate because it is located in a particular chromosome region suspected of being involved in the control of the trait, or its protein product may suggest that it could be involved in controlling the trait (e.g. milk protein genes in milk protein production). DNA: the genetic information in a genome is encoded in deoxyribonucleic acid (DNA), which is stored in the nucleus of a cell. DNA has two strands structured in a double helix, which is made of a sugar (deoxiribose), phosphate, and four chemical bases – the nucleotides: adenine (A), guanine (G), cytosine (C) and thymine (T). An A on one strand always pairs with a T on the other through two hydrogen bonds, while a C always pairs with a G through three hydrogen bonds. The two strands are, therefore, complementary to each other. Complementary DNA (cDNA): DNA sequences generated from the reverse transcription of mRNA sequences. This type of DNA includes exons and untranslated regions at the 5’ and 3’ ends of genes, but does not include intron DNA. Genetic marker: a DNA polymorphism that can be easily detected by molecular or phenotypic analysis. The marker can be within a gene or in DNA with no known function. Because DNA segments that lie near each other on a chromosome tend to be inherited together, markers are often used as indirect ways of tracking the inheritance pattern of a gene that has not yet been identified, but whose approximate location is known. Haplotype: a contraction of the phrase “haploid genotype”, is the genetic constitution of an individual chromosome. In the case of diploid organisms, the haplotype will contain one member of the pair of alleles for each site. It may refer to a set of markers (e.g. single nucleotide polymorphisms – SNPs) found to be statistically associated on a single chromosome.

374

With this knowledge, it is thought that the identification of a few alleles of a haplotype block can unambiguously identify all other polymorphic sites in this region. Such information is very valuable for investigating the genetics behind complex traits. Linkage: The association of genes and/or markers that lie near each other on a chromosome. Linked genes and markers tend to be inherited together. Linkage disequilibrium (LD): is a term used in the study of population genetics for the nonrandom association of alleles at two or more loci, not necessarily on the same chromosome. It is not the same as linkage, which describes the association of two or more loci on a chromosome with limited recombination between them. LD describes a situation in which some combinations of alleles or genetic markers occur more or less frequently in a population than would be expected from a random formation of haplotypes from alleles based on their frequencies. Linkage disequilibrium is caused by fitness interactions between genes or by such non-adaptive processes as population structure, inbreeding, and stochastic effects. In population genetics, linkage disequilibrium is said to characterize the haplotype distribution at two or more loci. Microarray technology: a new way of studying how large numbers of genes interact with each other and how a cell’s regulatory networks control vast batteries of genes simultaneously. The method uses a robot to precisely apply tiny droplets containing functional DNA to glass slides. Researchers then attach fluorescent labels to mRNA or cDNA from the cell they are studying. The labelled probes are allowed to bind to cDNA strands on the slides. The slides are put into a scanning microscope that can measure the brightness of each fluorescent dot; brightness reveals how much of a specific mRNA is present, an indicator of how active it is. Primer: a short (single strand) oligonucleotide sequence used in a polymerase chain reaction (PCR) RNA: Ribonucleic acid is a single stranded nucleic acid consisting of three of the four bases present in DNA (A, C and G). T is, however, replaced by uracil (U).

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

References Aebersold, R. & Mann, M. 2003. Mass spectrometrybased proteomics. Nature, 422 (6928): 198–207. Review. Ajmone-Marsan, P., Negrini, R., Milanesi, E., Bozzi, R., Nijman, I.J., Buntjer, J.B., Valentini, A. & Lenstra, J.A. 2002. Genetic distances within and across cattle breeds as indicated by biallelic AFLP markers. Animal Genetics, 33: 280–286. Akey, J.M., Zhang, G., Zhang, K., Jin, L. & Shriver, M.D. 2002. Interrogating a high-density SNP map for signatures of natural selection. Genome Research, 12(12): 1805–14. Aravin, A. & Tuschl, T. 2005. Identification and characterization of small RNAs involved in RNA silencing. Febs Letters, 579(26): 5830–40. Bachem, C.W.B., Van der Hoeven, R.S., De Bruijn, S.M., Vreugdenhil, D., Zabeau, M. & Visser, R.G.F. 1996. Visualization of differential gene expression using a novel method of RNA fingerprinting based on AFLP: analyses of gene expression during potato tuber development. The Plant Journal, 9: 745–753. Bamshad, M. & Wooding, S.P. 2003. Signatures of natural selection in the human genome. Nature Reviews Genetics, 4(2): 99–111. Review. Baumung, R., Simianer, H. & Hoffmann, I. 2004. Genetic diversity studies in farm animals – a survey, Journal of Animal Breeding and Genetics, 121: 361–373. Beaumont, M.A. & Balding, D.J. 2004. Identifying adaptive genetic divergence among populations from genome scans. Molecular Ecology, 13(4): 969–80. Beja-Pereira, A., Alexandrino, P., Bessa, I., Carretero, Y., Dunner, S., Ferrand, N., Jordana, J., Laloe, D., Moazami-Goudarzi, K., Sanchez, A. & Cañon, J. 2003. Genetic characterization of southwestern European bovine breeds: a historical and biogeographical reassessment with a set of 16 microsatellites. Journal of Heredity, 94: 243–50.

Berthier, D., Quere, R., Thevenon, S., Belemsaga, D., Piquemal, D., Marti, J. & Maillard, J.C. 2003. Serial analysis of gene expression (SAGE) in bovine trypanotolerance: preliminary results. Genetics Selection Evolution, 35 (Suppl. 1): S35–47. Bertone, P, Stolc, V., Royce, T.E., Rozowsky, J.S., Urban, A.E., Zhu, X., Rinn, J.L., Tongprasit, W., Samanta, M., Weissman, S., Gerstein, M. & Snyder, M. 2004. Global identification of human transcribed sequences with genome tiling arrays. Science, 306: 2242–2246. Black, W.C., Baer, C.F., Antolin, M.F. & DuTeau, N.M. 2001. Population genomics: genome-wide sampling of insect populations. Annual Review of Entomology, 46: 441–469. Bruford, M.W., Bradley, D.G. & Luikart, G. 2003. DNA markers reveal the complexity of livestock domestication. Nature Reviews Genetics, 4: 900–910. Buntjer, J.B., Otsen, M., Nijman, I.J., Kuiper, M.T. & Lenstra, J.A. 2002. Phylogeny of bovine species based on AFLP fingerprinting. Heredity, 88: 46–51. Campbell, D. & Bernatchez, L. 2004. Generic scan using AFLP markers as a means to assess the role of directional selection in the divergence of sympatric whitefish ecotypes. Molecular Biology and Evolution, 21(5): 945–56. Cañon, J., Garcıa, D., Garcıa-Atance, M.A., ObexerRuff, G., Lenstra, J.A., Ajmone-Marsan, P., Dunner, S. & The ECONOGENE Consortium. 2006. Geographical partitioning of goat diversity in Europe and the Middle East. Animal Genetics, 37: 327–334. Chen, S.Y., Su, Y.H., Wu, S.F., Sha, T. & Zhang, Y.P. 2005. Mitochondrial diversity and phylogeographic structure of Chinese domestic goats. Molecular Phylogenetics and Evolution, 37: 804–814. Clark, A.G., Hubisz, M.J., Bustamante, C.D., Williamson, S.H. & Nielsen, R. 2005. Ascertainment bias in studies of human genomewide polymorphism. Genome Research, 15: 1496–1502.

375

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

Clop, A., Marcq, F., Takeda, H., Pirottin, D., Tordoir, X., Bibe, B., Bouix, J., Caiment, F., Elsen, J.M., Eychenne, F., Larzul, C., Laville, E., Meish, F., Milenkovic, D., Tobin, J., Charlier, C. & Georges, M. 2006. A mutation creating a potential illegitimate microRNA target site in the myostatin gene affects muscularity in sheep. Nature Genetics, 38: 813–818. De Marchi, M., Dalvit, C., Targhetta, C. & Cassandro, M. 2006. Assessing genetic diversity in indigenous Veneto chicken breeds using AFLP markers. Animal Genetics, 37: 101–105. Donson, J., Fang, Y., Espiritu-Santo, G., Xing, W., Salazar, A., Miyamoto, S., Armendarez, V. & Volkmuth, W. 2002. Comprehensive gene expression analysis by transcript profiling. Plant Molecular Biology, 48: 75–97. Excoffier, L., Smouse, P.E. & Quattro, J.M. 1992 Analysis of molecular variance inferred from metric distances among DNA haplotypes: application to human mitochondrial DNA restriction data. Genetics, 131: 479–491. Farnir, F., Grisart, B., Coppieters, W., Riquet, J., Berzi, P., Cambisano, N., Karim, L., Mni, M., Moisio, S., Simon, P., Wagenaar, D., Vilkki, J. & Georges, M. 2002. Simultaneous mining of linkage and linkage disequilibrium to fine map quantitative trait loci in outbred half-sib pedigrees: revisiting the location of a quantitative trait locus with major effect on milk production on bovine chromosome 14. Genetics, 161: 275–287. Fields, S. & Song, O. 1989. A novel genetic system to detect protein–protein interactions. Nature, 340: 245–246. Freeman, A.R., Bradley, D.G., Nagda, S., Gibson, J.P. & Hanotte, O. 2006. Combination of multiple microsatellite data sets to investigate genetic diversity and admixture of domestic cattle. Animal Genetics, 37: 1–9. Goldstein, D.B., Linares, A.R., Cavalli-Sforza, L.L. & Feldman, M.W. 1995. An evaluation of genetic distances for use with microsatellite loci. Genetics, 139: 463–471.

376

Goldstein, D.B. & Schlötterer, C. 1999. Microsatellites: evolution and applications. New York. Oxford University Press. Grisart, B., Coppieters, W., Farnir, F., Karim, L., Ford, C., Berzi, P., Cambisano, N., Mni, M., Reid, S., Simon, P., Spelman, R., Georges, M. & Snell, R. 2002. Positional candidate cloning of a QTL in dairy cattle: identification of a missense mutation in the bovine DGAT1 gene with major effect on milk yield and composition. Genome Research, 12: 222–231. Guo, J., Du, L.X., Ma, Y.H., Guan, W.J., Li, H.B., Zhao, Q.J., Li, X. & Rao, S.Q. 2005. A novel maternal lineage revealed in sheep (Ovis aries). Animal Genetics, 36: 331–336. Haley, C. & de Koning, D.J. 2006. Genetical genomics in livestock: potentials and pitfalls. Animal Genetics, 37(Suppl 1): 10–12. Hanotte, O., Bradley, D.G., Ochieng, J.W., Verjee, Y. & Hill, E.W. 2002. African pastoralism: genetic imprints of origins and migrations. Science, 296: 336–339. Hayes, B.J., Visscher, P.M., McPartlan, H.C. & Goddard, M.E. 2003. A novel multilocus measure of linkage disequilibrium to estimate past effective population size. Genome Research, 13: 635–643. Hill, E.W., O’Gorman, G.M., Agaba, M., Gibson, J.P., Hanotte, O., Kemp, S.J., Naessens, J., Coussens, P.M. & MacHugh, D.E. 2005. Understanding bovine trypanosomiasis and trypanotolerance: the promise of functional genomics. Veterinary Immunology and Immunopathology, 105: 247–258. Hill, W.G. 1981. Estimation of effective population size from data on linkage disequilibrium. Genetics Research, 38: 209–216. Hillel, J., Groenen, M.A., Tixier-Boichard, M., Korol, A.B., David, L., Kirzhner, V.M., Burke, T., BarreDirie, A., Crooijmans, R.P., Elo, K., Feldman, M.W., Freidlin, P.J., Maki-Tanila, A., Oortwijn, M., Thomson, P., Vignal, A., Wimmers, K. & Weigend, S. 2003. Biodiversity of 52 chicken populations assessed by microsatellite typing of DNA pools. Genetics Selection Evolution, 35: 533–557.

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

Hood, L., Heath, J.R., Phelps, M.E. & Lin, B. 2004. Systems biology and new technologies enable predictive and preventative medicine. Science, 306: 640–643. Ibeagha-Awemu, E.M., Jann, O.C., Weimann, C. & Erhardt, G. 2004. Genetic diversity, introgression and relationships among West/Central African cattle breeds. Genetics Selection Evolution, 36: 673–690. Jarne, P. & Lagoda, P.J.L. 1996. Microsatellites, from molecules to populations and back. Tree, 11: 424–429. Joshi, M.B., Rout, P.K., Mandal, A.K., Tyler-Smith, C., Singh, L. & Thangaraj, K. 2004. Phylogeography and origin of Indian domestic goats. Molecular Biology and Evolution, 21: 454–462. Joost, S. 2006. The geographical dimension of genetic diversity. A GIScience contribution for the conservation of animal genetic resources. École Polytechnique Fédérale de Lausanne, Switzerland. (PhD thesis) Kayser, M., Brauer, S. & Stoneking, M. 2003. A genome scan to detect candidate regions influenced by local natural selection in human populations. Molecular Biology and Evolution, 20: 893–900. Lai, S.J., Liu, Y.P., Liu, Y.X., Li, X.W. & Yao, Y.G. 2006. Genetic diversity and origin of Chinese cattle revealed by mtDNA D-loop sequence variation. Molecular Phylogenetics and Evolution, 38: 146–54. Lan, L., Chen, M., Flowers, J.B., Yandell, B.S., Stapleton, D.S., Mata, C.M., Ton-Keen Mui, E., Flowers, M.T., Schueler, K.L., Manly, K.F., Williams, R.W., Kendziorski, C. & Attie, A.D. 2006. Combined expression trait correlations and expression quantitative trait locus mapping. PLoS Genetics, 2: 51–61. Liang, P. & Pardee, A.B. 1992. Differential display of eukaryotic messenger RNA by means of the polymerase chain reaction. Science, 257: 967–997.

Liu, Y.P., Wu, G.S., Yao, Y.G., Miao, Y.W., Luikart, G., Baig, M., Beja-Pereira, A., Ding, Z.L., Palanichamy, M.G. & Zhang, Y.P. 2006. Multiple maternal origins of chickens: out of the Asian jungles. Molecular Phylogenetics and Evolution, 38: 12–19. Lueking, A., Possling, A., Huber, O., Beveridge, A., Horn, M., Eickhoff, H., Schuchardt, J., Lehrach, H. & Cahill, D.J. 2003. A nonredundant human protein chip for antibody screening and serum profiling. Molecular and Cellular Proteomics, 2: 1342–1349. Luikart, G., England, P.R., Tallmon, D., Jordan, S. & Taberlet, P. 2003. The power and promise of population genomics: from genotyping to genome typing. Nature Reviews Genetics, 4: 981–994. Luikart, G., Gielly, L., Excoffier, L., Vigne, J.D., Bouvet, J. & Taberlet, P. 2001. Multiple maternal origins and weak phylogeographic structure in domestic goats. Proceedings of the National Academy of Science USA, 98: 5927–5932. Mburu, D.N., Ochieng, J.W., Kuria, S.G., Jianlin, H. & Kaufmann, B. 2003. Genetic diversity and relationships of indigenous Kenyan camel (Camelus dromedarius) populations: implications for their classification. Animal Genetics, 34(1): 26–32. McPherron, A.C. & Lee, S.J. 1997. Double muscling in cattle due to mutations in the myostatin gene. Proceedings of the National Academy of Science USA, 94: 12457–12461. Negrini, R., Milanesi, E., Bozzi, R., Pellecchia, M. & Ajmone-Marsan, P. 2006. Tuscany autochthonous cattle breeds: an original genetic resource investigated by AFLP markers. Journal of Animal Breeding and Genetics, 123: 10–16. Nei, M. 1972. Genetic distance between populations. The American Naturalist, 106: 283–292. Nei, M. 1978. Estimation of average heterozygosity and genetic distance from a small number of individuals. Genetics, 89: 583–590. Nei, M. & Roychoudhury, A.K. 1974. Sampling variances of heterozygosity and genetic distance. Genetics, 76: 379–390.

377

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

Nei, M., Tajima, F. & Tateno, Y. 1983. Accuracy of estimated phylogenetic trees from molecular data. II. Gene frequency data. Journal of Molecular Evolution, 19: 153–170. Nielsen, R. & Signorovitch, J. 2003. Correcting for ascertainment biases when analyzing SNP data: applications to the estimation of linkage disequilibrium. Theoretical Population Biology, 63: 245–55. Nijman, I.J., Otsen, M., Verkaar, E.L., de Ruijter, C. & Hanekamp, E. 2003. Hybridization of banteng (Bos javanicus) and zebu (Bos indicus) revealed by mitochondrial DNA, satellite DNA, AFLP and microsatellites. Heredity, 90: 10–16. Pariset, L., Cappuccio, I., Joost, S., D’Andrea, M.S., Marletta, D., Ajmone Marsan, P., Valentini A. & ECONOGENE Consortium 2006. Characterization of single nucleotide polymorphisms in sheep and their variation as an evidence of selection. Animal Genetics, 37: 290–292. Pritchard, J.K., Stephens, M. & Donnelly, P. 2000. Inference of population structure using multilocus genotype data. Genetics, 155: 945–959. Rabie, T.S., Crooijmans, R.P., Bovenhuis, H., Vereijken, A.L., Veenendaal, T., van der Poel, J.J., Van Arendonk, J.A., Pakdel, A. & Groenen, M.A. 2005. Genetic mapping of quantitative trait loci affecting susceptibility in chicken to develop pulmonary hypertension syndrome. Animal Genetics, 36: 468–476. Saitou, N. & Nei, M. 1987. The neighbor-joining method: a new method for reconstructing phylogenetic trees. Molecular Biology and Evolution, 4: 406–425. Sachidanandam, R., Weissman, D., Schmidt, S.C., Kakol, J.M., Stein, L.D., Marth, G., Sherry, S., Mullikin, J.C., Mortimore, B.J., Willey, D.L., Hunt, S.E., Cole, C.G., Coggill, P.C., Rice, C.M., Ning, Z., Rogers, J., Bentley, D.R., Kwok, P.Y., Mardis, E.R., Yeh, R.T., Schultz, B., Cook, L., Davenport, R., Dante, M., Fulton, L., Hillier, L., Waterston, R.H., McPherson, J.D., Gilman, B., Schaffner, S., Van Etten, W.J., Reich, D., Higgins, J., Daly, M.J., Blumenstiel, B., Baldwin, J., StangeThomann, N., Zody, M.C., Linton, L., Lander, E.S. & Altshuler, D.; International SNP Map Working

378

Group. 2001. A map of human genome sequence variation containing 1.42 million single nucleotide polymorphisms. Nature, 409: 928–933. SanCristobal, M., Chevalet, C., Haley, C.S., Joosten, R., Rattink, A.P., Harlizius, B., Groenen, M.A., Amigues, Y., Boscher, M.Y., Russell, G., Law, A., Davoli, R., Russo, V., Desautes, C., Alderson, L., Fimland, E., Bagga, M., Delgado, J.V., VegaPla, J.L., Martinez, A.M., Ramos, M., Glodek, P., Meyer, J.N., Gandini, G.C., Matassino, D., Plastow, G.S., Siggens, K.W., Laval, G., Archibald, A.L., Milan, D., Hammond, K. & Cardellino, R. 2006a. Genetic diversity within and between European pig breeds using microsatellite markers. Animal Genetics, 37: 189–198. SanCristobal, M., Chevalet, C., Peleman, J., Heuven, H., Brugmans, B., van Schriek, M., Joosten, R., Rattink, A.P., Harlizius, B., Groenen, M.A., Amigues, Y., Boscher, M.Y., Russell, G., Law, A., Davoli, R., Russo, V., Desautes, C., Alderson, L., Fimland, E., Bagga, M., Delgado, J.V., Vega-Pla, J.L., Martinez, A.M., Ramos, M., Glodek, P., Meyer, J.N., Gandini, G., Matassino, D., Siggens, K., Laval, G., Archibald, A., Milan, D., Hammond, K., Cardellino, R., Haley, C. & Plastow, G. 2006b. Genetic diversity in European pigs utilizing amplified fragment length polymorphism markers. Animal Genetics, 37: 232–238. Sauer, S., Lange, B.M.H., Gobom, J., Nyarsik, L., Seitz, H. & Lehrach, H. 2005. Miniaturization in functional genomics and proteomics. Nature Reviews Genetics, 6: 465–476. Sodhi, M., Mukesh, M., Mishra, B.P., Mitkari, K.R., Prakash, B. & Ahlawat, S.P. 2005. Evaluation of genetic differentiation in Bos indicus cattle breeds from Marathwada region of India using microsatellite polymorphism. Animal Biotechnology, 16: 127–137. Storz, G., Altuvia, S. & Wassarman, K.M. 2005. An abundance of RNA regulators. Annual Review of Biochemistry, 74: 199–217. Sunnucks, P. 2001. Efficient genetic markers for population biology. Tree, 15: 199–203.

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

Syvänen, A.C. 2001. Accessing genetic variation genotyping single nucleotide polymorphisms. Nature Reviews Genetics, 2: 930–941.

Wienholds, E. & Plasterk, R.H. 2005. MicroRNA function in animal development. FEBS Letters, 579: 5911–5922.

Takezaki, N. & Nei, M. 1996. Genetic distances and reconstruction of phylogenetic trees from microsatellite DNA. Genetics, 144: 389–399.

Wong, G.K., Liu, B., Wang, J., Zhang, Y., Yang, X., Zhang, Z., Meng, Q., Zhou, J., Li, D., Zhang, J., Ni, P., Li, S., Ran, L., Li, H., Zhang, J., Li, R., Li, S., Zheng, H., Lin, W., Li, G., Wang, X., Zhao, W., Li, J., Ye, C., Dai, M., Ruan, J., Zhou, Y., Li, Y., He, X., Zhang, Y., Wang, J., Huang, X., Tong, W., Chen, J., Ye, J., Chen, C., Wei, N., Li, G., Dong, L., Lan, F., Sun, Y., Zhang, Z., Yang, Z., Yu, Y., Huang, Y., He, D., Xi, Y., Wei, D., Qi, Q., Li, W., Shi, J., Wang, M., Xie, F., Wang, J., Zhang, X., Wang, P., Zhao, Y., Li, N., Yang, N., Dong, W., Hu, S., Zeng, C., Zheng, W., Hao, B., Hillier, L.W., Yang, S.P., Warren, W.C., Wilson, R.K., Brandstrom, M., Ellegren, H., Crooijmans, R.P., van der Poel, J.J., Bovenhuis, H., Groenen, M.A., Ovcharenko, I., Gordon, L., Stubbs, L., Lucas, S., Glavina, T., Aerts, A., Kaiser, P., Rothwell, L., Young, J.R., Rogers, S., Walker, B.A., van Hateren, A., Kaufman, J., Bumstead, N., Lamont, S.J., Zhou, H., Hocking, P.M., Morrice, D., de Koning, D.J., Law, A., Bartley, N., Burt, D.W., Hunt, H., Cheng, H.H., Gunnarsson, U., Wahlberg, P., Andersson, L., Kindlund, E., Tammi, M.T., Andersson, B., Webber, C., Ponting, C.P., Overton, I.M., Boardman, P.E., Tang, H., Hubbard, S.J., Wilson, S.A., Yu, J., Wang, J., Yang, H.; International Chicken Polymorphism Map Consortium. 2004. A genetic variation map for chicken with 2.8 million single-nucleotide polymorphisms. Nature, 432: 717–722.

Tapio, M., Tapio, I., Grislis, Z., Holm, L.E., Jeppsson, S., Kantanen, J., Miceikiene, I., Olsaker, I., Viinalass, H. & Eythorsdottir, E. 2005. Native breeds demonstrate high contributions to the molecular variation in northern European sheep. Molecular Ecology, 14: 3951–3963. Tilquin, P., Barrow, P.A., Marly, J., Pitel, F., PlissonPetit, F., Velge, P., Vignal, A., Baret, P.V., Bumstead, N. & Beaumont, C. 2005. A genome scan for quantitative trait loci affecting the Salmonella carrier-state in the chicken. Genetics Selection Evolution, 37: 539–61. Troy, C.S., MacHugh, D., Bailey, J.F., Magee, D.A., Loftus, R.T., Cunningham, P., Chamberlain, A.T., Sykesk, B.C. & Bradley D.G. 2001. Genetic evidence for Near-Eastern origins of European cattle. Nature, 410: 1088–1091. Velculescu, V.E., Vogelstein, B. & Kinzler, K.W. 2000. Analyzing uncharted transcriptomes with SAGE. Trends in Genetics, 16: 423–425. Velculescu, V.E., Zhang, L., Vogelstein, B. & Kinzler, K.W. 1995. Serial analysis of gene expression. Science, 270: 484–487. Vos, P., Hogers, R., Bleeker, M., Reijans, M., van de Lee, T., Hornes, M., Frijters, A., Pot, J., Peleman, J. & Kuiper, M. 1995. AFLP: a new technique for DNA fingerprinting. Nucleic Acids Research, 23: 4407–1444.

Zhu, H., Bilgin, M. & Snyder, M. 2003. Proteomics. Annual Review of Biochemistry, 72: 783–812.

Weir, B.S. & Basten, C.J. 1990. Sampling strategies for distances between DNA sequences. Biometrics, 46: 551–582. Weir, B.S. & Cockerham, C.C. 1984. Estimating Fstatistics for the analysis of population structure. Evolution, 38: 1358–1370.

379

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

Section D

Genetic improvement methods to support sustainable utilization 1

Introduction

This section gives an overview of genetic improvement methods for sustainable use of AnGR. The first chapter describes the contexts for genetic improvement. As social and economic contexts are discussed extensively in other parts of the Report, they are only briefly described here. The scientific and technology-related context is described in greater detail. The second chapter discusses breeding strategies for genetic improvement, along with the elements of a straight-breeding programme. These elements involve planning, implementation and evaluation, and constitute a continuous and interactive process. Breeding programmes for the main livestock species in high-input systems are then reviewed. This includes a description not only of the breeding goals and the traits making up the selection criteria, but also the organization and the evolution of the breeding sector. This is followed by a description of breeding strategies for low-input systems, and those utilized in the context of breed conservation. This distinction is somewhat artificial as the situations and strategies sometimes overlap. Finally, some general conclusions are drawn.

2

The context for genetic improvement

Genetic improvement implies change. For a change to be an improvement, the overall effects of the change must bring positive benefits

to the owners of the animals in question or to the owners’ community. Moreover, to be an improvement, the effects of the change should bring positive benefits in both the short and the long term, or at minimum a short-term benefit should not result in long-term harm. As such, it is vital that the planning of genetic improvement programmes takes careful account of the social, economic and environmental context in which they will operate. This can best be achieved by making these programmes an integral part of national livestock development plans, which should establish broad development objectives for each production environment.

2.1 Changing demand Traditionally, livestock breeding has been of interest only to a small number of professionals: breeding company employees, farmers, and some animal scientists. However, food production is changing from being producer driven to consumer driven. Consumer confidence in the livestock industry has broken down in many countries (Lamb, 2001). Fears about the quality and safety, of animal products have been heightened in recent years by various crises: bovine spongiform encephalopathy (BSE), dioxin, and more recently, highly pathogenic avian influenza (HPAI). Welfare has also become an important element in consumers’ perception of product quality especially in Europe (organic products and freerange animals). At the same time, the majority of consumers have become less connected to the

381

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

countryside, and know less about farming. There is a growing demand for “natural” production, but often without a clear understanding of what this should encompass.

2.4 Scientific and technological advances Developments in genetic improvement methods

2.2 Diverse production environments Sustainable production systems need to be tailored to account for physical, social and market conditions. For breeding organizations this raises the question of whether they should diversify their breeding objectives, or whether they should breed an animal that can do well under a wide range of environments (physical environment, management system and market conditions). To date, however, only limited insights into the underlying genetics of phenotypic adaptation to the environment have been achieved.

2.3 Increasing recognition of the importance of genetic diversity Livestock breeding requires variability within and between populations if it is to improve the traits of interest. Genetic diversity is important to meet present requirements, but is especially important to meet future requirements. For example, a change of emphasis from high-input to low-input production systems will favour different breeds and different characteristics within breeds. More generally, the increasing importance given to factors such as animal welfare, environmental protection, distinctive product quality, human health and climate change, will require a wider range of criteria to be included in breeding programmes. These criteria are often met by local breeds. Thus, it is possible that the most appropriate strategies for managing these breeds may involve only limited genetic change. For example, it may be wise to maintain adaptation to the local environment and disease challenges – and even to maintain the level of a production trait, such as body size or milk production, if this is currently at or near an optimum level.

382

Quantitative genetics A breeding scheme aims to achieve genetic improvement in the breeding goal through the selection of the animals that will produce the next generation. The breeding goal reflects the traits that the breeder aims to improve through selection. The rate of genetic improvement ($G) with respect to the breeding goal (and the underlying traits) depends on the amount of genetic variability in the population, the accuracy of the selection criteria, the intensity of selection, and the generation interval. Maintenance of genetic variation is a condition for continuous genetic improvement. Genetic variation is lost by genetic drift and gained by mutation. Therefore, the minimum population size to maintain genetic variation is a function of the mutation rate (Hill, 2000). Selection experiments in laboratory animals have shown that substantial progress can be maintained for many generations, even in populations with an effective size well under 100, but that responses increase with population size (ibid.). The loss of genetic variation within a breed is related to the rate of inbreeding (ΔF). In the absence of selection, ΔF is related directly to the number of sires and dams. In populations undergoing selection, this assumption is no longer valid because parents contribute unequally to the next generation. A general theory to predict rates of inbreeding in populations undergoing selection has recently been developed (Woolliams et al., 1999; Woolliams and Bijma, 2000). This approach facilitates a deterministic optimization of short and long-term response in breeding schemes. Research on the optimization of breeding schemes initially focused on genetic gain, while little attention was paid to inbreeding. It is now well accepted that constraining inbreeding is

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

an important element of breeding schemes. Meuwissen (1997) developed a dynamic selection tool which maximizes genetic gain while restricting the rate of inbreeding. From a given set of selection candidates, the method allows the selection of a group of parents in which the genetic merit is maximized while the average coefficient of coancestry is constrained. Implementation of this method results in a dynamic breeding programme, in which the number of parents and the number of offspring per parent may vary, depending on the candidates available in a particular generation. The accuracy of selection depends largely on the quality and the quantity of the performance records that are available. Genetic improvement can only be made if performance and pedigree are recorded. Based on these observations, the genetic merit of an individual is predicted and the animals with the highest predicted merit can be selected as parents. It is well established that the method of choice for the genetic evaluation of linear traits (e.g. milk and egg production, body size and feed efficiency) is best linear unbiased prediction based on an animal model (BLUP-AM) (Simianer, 1994). The development of algorithms and software has meant that by today, in most countries and for most species, BLUP-AM is routinely used by breeding companies or in national-level breeding programmes. The limitations associated with applying simplistic single-trait models has led to the development of multiple-trait BLUPAM evaluations based on sophisticated models (including, for example, maternal effects, herd × sire interactions or dominance genetic effects). This has been greatly facilitated by the increasing power of computers, and major advances in computational methods. The tendency now is to use all available information, including single test day records, records from cross-bred animals, and a wide geographical range (across countries). Significant difficulties associated with the use of increasingly complex models are a lack of robustness (especially when population size is limited) and computational problems. The challenge today is to develop tools to systematically validate the models used.

BLUP is optimum only when the true genetic parameters are known. Methods for unbiased estimation of (heterogeneous) variance components with large data sets have been developed. Restricted Maximum Likelihood (REML) applied to animal models is the method of preference. Quite a few important traits are not correctly described by linear models (e.g. traits based on scoring and survival). A wide variety of nonlinear mixed models have, therefore, been proposed: threshold models, survival models, models based on ranks, Poisson models, etc. However, the benefits of using these nonlinear models remain to be proven. The selection intensity reflects the proportion of animals that are needed as parents for the next generation. Reproductive capacity and techniques have an important influence on the number of parents that are needed for the production of the next generation, and thereby on the rate of genetic improvement. In poultry, high reproductive capacity means that about 2 and 10 percent of the male and female candidates, respectively, are retained as parents. In cattle, the introduction of AI has resulted in an enormous reduction in the number of sires. In dairy and beef cattle, the bulls used for AI and the cows with high genetic merit are the nucleus animals, and form less than 1 percent of the entire population. The generation interval is the average time between two generations. In most populations, a number of age classes can be distinguished. The amount of information available differs between classes. In general, there is less information about the younger age classes than about older age classes. Consequently, the accuracy of estimates of breeding value is lower in the younger generations. However, the mean level of the estimated breeding value (EBV) of young age classes is higher than that of older age classes because of continuous genetic improvement in the population. Selection across age classes to obtain the highest selection differential is recommended (James, 1972). The fraction of animals selected from each age class depends on the differences

383

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

in accuracy of the EBV between the age classes (Ducrocq and Quaas, 1988; Bijma et al., 2001). The use of reproductive technologies may increase the amount of sib information available, and thereby increase the accuracy of the EBV of younger age classes (van Arendonk and Bijma, 2003). This will change the proportion of parents selected from the younger age classes, and therefore also influence the average generation interval. Thus, generation interval is primarily a result of selection among the available age classes.

Molecular genetics Molecular genetics in livestock has been subject to extensive study during the last two decades. These studies are related to gene-based selection of Mendelian traits (mainly diseases and genetic defects), marker assisted selection and introgression. Furthermore, molecular information is increasingly used to assist breed conservation programmes and to improve understanding of the origin and domestication of livestock. Gene-based selection. Increasing knowledge of the animals’ genome increases the prospects for applying this technology and provides new tools with which to select for healthy animals. Initial applications are related to Mendelian traits. In cattle for example, DNA diagnosis is routinely utilized to eliminate genetic disorders such as bovine leukocyte adhesion deficiency (BLAD), deficiency of uridine monophosphate synthase (DUMPS) and complex vertebral malformation (CVM), as well as in selection for traits such as milk kappa-casein and double muscling. In pigs, the best-known gene which has so far been used in commercial breeding is the “halothane” gene. It was known that a number of pigs could not handle stressful situations (e.g. transportation to the slaughterhouse). A (recessive) gene – a natural mutation, called the “halothane” gene – was found to be responsible for this defect. Using a DNA test that detects whether a pig has the “defective form” of the gene, it has been possible to eliminate this gene completely from several breeds (Fuji et al., 1991).

384

Scrapie, the prion disease of sheep, is the most common natural form of transmissible spongiform encephalopathy (TSE), a group of diseases which also include Creutzfeldt-Jakob disease in humans and BSE in cattle. Genetic susceptibility to scrapie is strongly modulated by allelic variations at three different codons in the sheep PrP gene (Hunter, 1997). Breeding for scrapie resistance has, therefore, been considered an attractive option for the control of this disease (Dawson et al., 1998; Smits et al., 2000). This can be done by selecting for the allele that is associated with the greatest degree of resistance to scrapie (the ARR allele). As described in Part 1 – Section F: 4, breeding programmes to eliminate scrapie can pose a threat to rare breeds that have a low frequency of the resistant genotype. Marker assisted selection. Most economically important traits in animal production are of a quantitative nature and are affected by a large number of genes (loci), a few of which have major effects, while the majority have small effects (Le Roy et al., 1990; Andersson et al., 1994). If a gene (locus) with a major effect can be identified, and if a molecular test can be designed, animals’ genotypes at the locus can be used for selection. In other cases, a chromosomal region close to the gene of interest may be identified and used as a marker. Mixed models of inheritance, which assume one or several identified segregating loci, and an additional polygenic component, have been developed. When genotypes at each identified locus are known, they can be treated as fixed effects in standard mixed-model techniques (Kennedy et al., 1992). When only genotypes at linked markers are known, the uncertainty resulting from unknown haplotypes and recombination events has to be taken into account (Fernando and Grossman, 1989). Extra genetic gain is usually to be expected if information on genes with medium to large effects is included in the genetic evaluation process. Numerous studies have investigated this problem in recent years. Results are not always comparable, because selection criteria differed

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

between studies (i.e. from an index based on individual information to animal models), but they all indicate that knowledge of genotypes at quantitative trait loci generally improves short-term response to selection (Larzul et al., 1997). Conversely, some discrepancies have been obtained for long-term response to selection – see Larzul et al. (1997). In less favourable situations where only genotypes at linked markers are known, results largely depend on the particular circumstances. Large gains can be expected when linkage disequilibrium exists at the population level (Lande and Thompson, 1990), and when traits are difficult to measure (e.g. disease resistance), sex limited (e.g. traits related to egg or milk production), expressed late in the lifespan of the animals (e.g. longevity and persistency in litter size), or measured after slaughtering (e.g. meat quality traits). In other cases, the advantage of marker assisted selection may be questionable. Genes at the same or at different loci interact with each other in producing a phenotypic effect. It is seldom known how this occurs. When, by using statistical models, an apparent effect is assigned to a particular gene, such interaction is not taken into account. This explains, at least partly, why even when genes with major effects are identified, incorporating them (or their markers) into a selection programme may not achieve the desired results. Because of such interactions, there is often an apparent lack of consistency between different studies related to the use of genetic markers (Rocha et al., 1998). To correctly assess the effect of a gene, the average effect over the possible genotypes in the population where the information is to be applied (weighted according to their frequencies) has to be considered. Introgression is advocated mainly to improve disease resistance in a given population. If markers for the resistance gene(s) (or probe for the gene) are available, marker assisted selection may be used to simplify the process of introgression. Dekkers and Hospital (2002) discuss the use of repeated backcrosses to introgress a gene into a population. If the non-resistant breed is considered the recipient breed, and the breed

that carries the resistance gene is considered the donor breed, introgression of the desirable gene from the donor breed to the recipient breed is accomplished by multiple backcrosses to the recipient breed, followed by one or more generations of intercrossing. The aim of the backcross generations is to generate individuals that carry one copy of the donor gene, but that are similar to the recipient breed for the rest of the genome. The aim of the intercrossing phase is to fix the donor gene. Marker information can enhance the effectiveness of the backcrossing phase of gene introgression strategies by identifying carriers of the target gene (foreground selection), and by enhancing recovery of the recipient genetic background (background selection). Generally, it is more feasible and economically sound to mate, in successive generations, pure-bred females of the recipient breed to cross-bred males that carry the desired gene, than to carry out the reverse process. If the gene for resistance is dominant, its introgression into a population may be effective even without a molecular marker for the gene. If the gene for resistance is recessive (or codominant), markers are necessary. In cases where resistance is polygenic, introgression without genetic markers is not likely to be effective; by the time the genetic influence of the donor breed is high enough to give high levels of resistance, the desired characteristics of the recipient breed will probably have been lost. In fact, the development of a composite breed would be easier than the introgression of numerous genes into a recipient breed by backcrossing, even when genetic markers are available. Hanotte et al. (2003) mapped QTLs affecting trypanotolerance in a cross between the “tolerant” N’Dama and “non-tolerant” Boran cattle breeds. Results showed that at some of the putative QTLs associated with trypanotolerance, the allele associated with tolerance came from the non-tolerant cattle. It was concluded that “selection for trypanotolerance within an F2 cross between N’Dama and Boran cattle could produce a synthetic breed with higher trypanotolerance levels than currently exist in the parental breeds.”

385

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

Conceptually, introgression through marker assisted selection could be accomplished even without exposure to the disease agent. It is, however, wise to test the resistance of animals with the desired genotype. Molecular characterization of genetic diversity is helpful in the planning of conservation programmes and to develop understanding of the origin and domestication of livestock species. Better knowledge of genomic variation, together with the development of new quantitative genetic methods, may provide the means to link marker information to functional variation. For example, combination of molecular methods and pedigree analysis has been used to estimate the degree of genetic diversity in founder populations in thoroughbred horses (Cunningham et al., 2001).

Developments in reproductive technologies Reproductive technology has a direct effect on the rate of genetic improvement. For a given population size, a higher reproduction rate implies a lower number of breeding animals and, therefore, a higher intensity of selection. More offspring per breeding animal also allows more accurate estimation of breeding values. Another advantage of increasing reproductive rates is to disseminate superior genetic stock more quickly. As reproductive technologies are extensively discussed elsewhere in the report, this chapter focuses only on the use of AI and multiple ovulation and embryo transfer (MOET) in breeding programmes. For other techniques, only a brief description is provided here. Artificial insemination. The use of AI results in higher selection intensity, more accurate selection of males based on progeny testing and more accurate estimation of breeding value across herds. The latter is a result of exchange of semen between different nucleus herds, which facilitates the establishment of genetic links between them. AI is used by breeding organizations for most species. For species such as cattle that have low reproductive rates, progeny testing based on AI is a prerequisite for an accurate estimation of breeding values for traits of low heritability such as

386

functional traits. AI allows faster dissemination of genetic superiority to the commercial population. Sixty to eighty percent of all the AI performed is carried out in cattle. A male identified as superior can leave thousands of progeny in different populations all over the world. AI requires technical skills both at the AI centre and on the farm, as well as effective lines of communication between the two. However, in many countries, the majority of producers are smallholder farmers, and existing skills and infrastructure may be insufficient to allow the successful operation of AI services. The farmer has to be able to detect heat and have a means to contact the semen distribution centre, which then has to be able to serve within few hours. For extensive production systems, this is a labourintensive process. Consequently, AI is unlikely to be used in extensive grazing systems for beef production. Similarly, AI is difficult to perform in sheep, and natural mating using superior males is still the dominant means of diffusing genetic improvement. Use of AI affects the ownership structure of the breeding sector. Where AI is used, the ownership of the breeding animals is usually transferred to larger breeding organizations, such as cooperatives or private breeding companies. For the last twenty years in the developed world, AI centres have been responsible for the identification of young bulls for progeny testing, and for the marketing of semen from proven sires. Multiple ovulation and embryo transfer. Increasing the reproductive rate of females by MOET is mainly useful in species with low reproductive rates such as cattle. The benefits are higher selection intensity on the female side, and more accurate estimation of breeding values. As family sizes are larger, there is more information available on animals’ sibs. This allows reasonably reliable breeding values to be obtained at a younger age, particularly when the traits are only recorded for one sex (female). In practice, this means that there is no need to wait for a progeny test to select males – they can be selected at younger age based on information on their

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

half-sib sisters. The gain in generation interval is large, and compensates for the loss of selection accuracy that results from replacing a progeny test by a sib test. The ability to select at a young age, even among embryos, is the main reason of the application of MOET in pig breeding. Embryo transfer is also used to disseminate desirable genes from superior female animals with minimum disease risks, as animals do not need to be transported. The use of MOET is costly and requires highly developed technical skills. The logistical challenge is that at the time of embryo transfer, a group of recipient cows needs to be available and synchronized. This can be done only in large centralized nucleus herds. In many cases, it may be better to invest resources in more basic prerequisites – performance and trait recording, extension and dissemination. This is all the more true as MOET seems less efficient than AI in enhancing genetic progress. In all cases, the introduction of AI and/or MOET has to be cost effective and accepted by the local farmers. Semen and embryo freezing gives breeding organizations the opportunity to create genebanks as a back-up store of genetic diversity in breeding programmes. Moreover, cryopreservation of gametes and embryos facilitates international exchange and transport of genetic material in ruminants, and is a prerequisite for routine use of AI and ET on a world scale. Cloning (somatic cells) is a new technology which is currently not being used commercially. This is partly for technical and economic reasons, and partly because there is no public desire for such developments at present. Cloning has potential application in the field of conservation, as other tissues may be easier to preserve than embryos. Sexing of embryos or semen enables the production of larger numbers of animals of a particular sex. For example, preferences for male or female offspring are obvious in cattle – females for milk production, and males for beef production. Numerous attempts have been made to develop a reliable technology. Currently, it is

possible to identify male and female embryos by various methods. However, with a few exceptions, this technology has not yet been widely used by breeders or farmers. Various attempts have been made to separate sperm based on their sexdetermining characteristics. However, further advances are required before the technology can be applied on a large scale. The use of the above-described reproductive and conservation techniques means that there is less need for the transportation of breeding animals. Furthermore, these technologies offer an opportunity to safeguard the health status of flocks and herds even when embryos originate from countries with a radically different health status.

2.5 Economic considerations Any economic evaluation should consider both returns and costs. As animal breeding is a longterm process, returns on breeding decisions may be realized many years later. This is the case in dairy cattle for example. Furthermore, different costs and returns are realized at different times with different probabilities, and a number of considerations that may not be important for relatively short-term processes are sometimes of major importance in the longer term. Until the advent of reproductive biotechnologies, the main cost elements of breeding programmes were trait measurement and recording, progeny testing and maintaining the breeding stock. Although the main objective of most recording systems is breeding, it should be noted that once available, the information is useful for other farm management decisions such as culling and predicting future production. Animal breeding in the developed world has become more and more sophisticated and professionalized, and hence costly. Economic considerations are, therefore, driving most if not all breeding-related activities, and economic theory has been incorporated into this area. The bases for economic evaluation are profit, economic efficiency, or return on investment. When breeding goals have been developed by

387

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

and for (groups of) producers, emphasis is put on profit maximization. In developing countries, markets are generally more local, but the same mechanism will apply. It is, therefore, advisable to opt for profit maximization, unless there are clear reasons to deviate from this strategy. A critical economic consideration is: who will pay for the genetic improvement? This question is not particularly important when breeding nuclei, multipliers and commercial herds/flocks are fully integrated. However, in all other situations, where vertical integration does not exist, it is not unusual that those who invested in breeding activities are unable to adequately recoup their investment. This commonly provides justification for public sector involvement in one or more facets of genetic improvement. Under a free market system, breeding organizations have to adapt to the demands of their customers – the commercial producers, who are normally only prepared to pay for improved breeding animals or semen if this will enhance their profits. However, it is interesting to note that even if a trend in breeding does not appear to be economically justified, it may continue for an extended period of time (Box 80). Under a government subsidized system, all or part of the costs of genetic efforts are paid for by taxpayers. In this case, breeding programmes should be subject to scrutiny to ensure that they truly produce some social benefits. Such benefits could include, for example, providing safer, more nutritious or less expensive products for the consumer, or reducing the negative environmental impacts of livestock production.

3

Elements of a breeding programme

The elements required in a breeding programme depend on the choice of the general breeding strategy. Thus, the first decision is which of the three main genetic improvement strategies should be applied: selection between breeds,

388

selection within breeds or lines, or cross-breeding (Simm, 1998). • Selection between breeds, the most radical option, is the substitution of a genetically inferior breed by a superior one. This can be done at once (when as in poultry the cost is not prohibitive) or gradually by repeated backcrossing with the superior breed (in large animals). • Cross-breeding, the second fastest method, capitalizes on heterosis and complementarity between breeds’ characteristics. Conventional cross-breeding systems (rotational systems and terminal sire-based systems) have been widely discussed (e.g. Gregory and Cundiff, 1980). The inter se mating of animals of newly developed composites has been suggested as an alternative form of cross-breeding (Dickerson, 1969; 1972). • The third method, within-breed selection, gives the slowest genetic improvement, especially if the generation interval is long. However, this improvement is permanent and cumulative, which is not the case for cross-breeding programmes. Gradual genetic improvement is the most sustainable form of improvement, as it gives the stakeholders time to adapt the production system to the intended change. When the traits of interest are numerous and/or some of them are antagonistic, different lines may be created, and maintained by within-line selection. These lines can then be crossed to produce commercial animals. This strategy is used in pig and poultry breeding. Setting up a breeding programme involves the definition of a breeding goal (Groen, 2000) and the design of a scheme that is able to deliver genetic progress in line with this goal. In practice, it involves the management of people and resources as well as the application of the principles of genetics and animal breeding (Falconer and Mackay, 1996). Each aspect of the breeding programme involves many processes, individuals and sometimes institutions. Success

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

depends on how well the available resources are harnessed and managed to achieve the goals of the stakeholders. The stakeholders of a breeding programme are all those who are affected, in one way or another, by its success. These include the end users of the products of the programme (i.e. livestock producers), commercial companies and others who directly or indirectly invest in the scheme, government departments, breed societies, and those employed to implement the programme. Other stakeholders include ancillary beneficiaries such as suppliers, distributors, and sellers of byproducts of the scheme. Most programmes have a pyramidal structure (Simm, 1998), with varying number of tiers depending on the sophistication of the programme. At the apex of the pyramid is the nucleus where selection and breeding of the elite pedigree animals is concentrated. The multiplication of stock happens in the middle tiers. This is required

when the number of nucleus animals is insufficient to satisfy the demands of commercial farmers. The bottom tier comprises the commercial units where the final product is disseminated. The pyramidal structure of the poultry breeding industry is illustrated in Figure 48. The activities that constitute a breeding programme can be summarized in eight major steps (Simm, 1998): • choice of breeding goal; • choice of selection criteria; • design of the breeding scheme; • recording of the animals; • genetic evaluation of the animals; • selection and breeding; • progress monitoring; and • dissemination of genetic improvement. These steps will be described in the following subchapters. However, the reader should be aware that planning, implementation and evaluation form a continuous process – the elements should

FIGURE 48 Structure of the poultry breeding industry

PRIMARY BREEDERS generating genetic progress

Grand parents / parents

MULTIPLIERS producing egg-type or meat type day-old chicks

EGG PRODUCER / BROILER GROWER

Final product (pullet / broiler)

EGG PROCESSING PLANT – SLAUGHTER HOUSE

DISTRIBUTION

CONSUMER

389

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

be approached interactively rather than stepby-step. A further critical element is the need to document in detail all areas of the breeding plan and its execution over time.

3.1 Breeding goals The breeding goal is a list of traits to be improved genetically. It should be in line with national agricultural development objectives, and appropriate for the production system for which it is defined and the breeds suited to the production system. A country’s development objectives for agricultural production traditionally include economic variables, but should be extended to accommodate ethics, and other social aspects of human well-being. These objectives are used to formulate the breeding goals. Different tools are available to achieve this. The most common is the profit function. In theory, setting up a profit function is straightforward, especially in the case of within-breed selection programmes, as it is a linear function of the relative economic values of the traits to be improved. In practice, however, it is not easy to obtain these economic values, partly because they may vary in time and in space, and partly because of a lack of time, expertise, knowledge, resources, etc. Thus, breeders manipulate the direction of change through trial and error based on perceived market demand and preference. Amer (2006) discusses other tools for formulating breeding goals such as the bioeconomic model and the geneflow model. Livestock improvement is measured relative to a given set of traits, generally referred to as “traits of economic importance”. In reality, the traits and their economic importance vary as widely as the breeding programmes. For many livestock species, the traits of economic importance are those that affect the productivity, longevity, health and reproductive ability of the animals. For most of the traits, the objective is a continuous improvement, but for some traits the goal is to reach intermediate values. Pharo and Pharo (2005) term these alternatives, respectively, breeding for a “direction” and for a “destination”.

390

An example of the latter is egg weight in laying hens. The market values eggs within a particular range of weights – for example, between 55 and 70 grams. Smaller eggs are not saleable and there is no premium for bigger ones. Given that egg size is correlated negatively to egg number, shell strength and hatchability, selecting for bigger eggs is not only a waste of selection intensity, it is also counter productive. Another example is body size. For meat animals, size at slaughter is an important determinant of value. Body size has a major effect on nutritional requirements, through its effect on maintenance requirements. It may also affect fertility. The latter (net fertility such as calf crop or lamb crop weaned) is a major determinant of biological efficiency and profitability. Since body size is associated with both costs and benefits, it is difficult to determine an optimum value, especially under grazing systems, because of the difficulty involved in adequately describing forage intake. Another consideration is that most slaughter markets discriminate against animals that fall outside a desired range of carcass (or live) weights. For example, the European market requires a minimum carcass weight, which cannot be met by some breeds (e.g. Sanga breeds from Namibia). Even if the current body size of these cattle is optimum with regard to biological efficiency, larger cattle may be more profitable. The choice of the breeding goal may be a oneoff activity, or one that is revised from time to time. The decision is taken by the breeders, with feedback from all tiers of the breeding pyramid. In poultry and pig breeding, this decision is taken by the top management of the breeding companies (research and development managers in agreement with technical and marketing or sales managers). In cattle breeding, the decision is taken at the apex nucleus, but usually in consultation with people in all other tiers including the commercial tier, in a way that reflects the ownership pattern of the programme. The outcome of breeding programmes, particularly in dairy and beef cattle, is realized many years after selection decisions are made.

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

Even in poultry, where the generation interval is shorter, a genetic change implemented in the nucleus will not be noticed at the commercial level in less than three years, at the earliest. This underlines the need to anticipate future demands when defining breeding goals. In a competitive market like the poultry breeding industry, the identification of traits of interest and the focus of selection efforts is not only highly dependent on signals from the market place (i.e. the commercial producers), but also on the performance of the products of competing programmes.

3.2 Selection criteria The breeding goal is distinct from the selection criteria that are used to take the decision as to which animals are to become the parents of the next generation. Usually, the decision involves the construction of a “selection index”. Measurements are taken in the candidate animals and their relatives, and are weighted according to index coefficients calculated to maximize the correlation between the selection index and the breeding goal. It should be emphasized that some of the breeding goal traits may differ from those used to construct the selection index. For example, pigs are selected for the fatness of their carcass – this is a breeding goal trait. However, it cannot be observed in selection candidates,

Box 80 Changing body size of beef cattle in the United States of America In 1900 the vast majority of beef cattle in the United States of America were Shorthorn, Hereford, or Angus. The cattle at the time were fairly large. Bulls of 1 100 kg and cows of 730 kg were common. Cattle were finished (fattened) primarily on grass, and there was some interest in producing cattle that would finish at a younger age and lighter weight. A trend developed for selecting for smaller-framed cattle that had greater apparent ability to fatten. Much of the selection was actually based on attempts to win in the show ring. Selection was effective, and major changes were achieved in the cattle population. After a few generations (the late 1920s and early 1930s) the cattle were probably of a more appropriate size for the production conditions under which they were kept. However, selection continued in the same direction, and by the 1950s the cattle in most highly regarded herds were much too small and predisposed to fattening to be profitable under any commercial management programme. A major change in the United States beef industry began in the mid-1950s, with the development of large feedlots in the Great Plains states. To be profitable in these new feedlots, cattle had to be able to grow at a fairly high rate for a long feeding

period (four or five months) without getting too fat. The small early fattening cattle which had previously been popular were not acceptable to the feedlot industry. Charolais and other continental European breeds became popular, and cattle of the British beef breeds were selected for increased size and growth. From the mid-1950s to the late 1960s, larger cattle were favoured as long as they were fairly compact in their conformation. However, by the late 1960s, larger cattle were favoured, even if they were taller and very different in their conformation from the popular cattle of the earlier period. Within a few years, cattle were being selected for larger frame size, even in the continental European breeds. This selection was also quite effective, and extremely large animals were produced. In the mid to late 1980s, several of the major breeding organizations realized that the trend had gone too far, and moves were made to produce more moderate sized animals. In the last ten years, more breeders have recognized that intermediate size is preferable to extremes in any direction. However, they continue to be in the minority, and extremely large cattle have continued to be favoured in many major herds.

391

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

as this would mean that they would have to be slaughtered. A predictor trait, the subcutaneous fat thickness measured ultrasonically, is therefore recorded. Where it is difficult or expensive to acquire information on the relationships between animals, and the traits are sufficiently heritable, selection can be based on individual performance (mass selection). The construction of the selection index is a technical issue, and requires personnel with the necessary expertise. There are numerous circumstances in which at the moment of selection many traits that are not relevant to breeding goal trait list are considered. This can seriously decrease the actual selection intensity and, therefore, limit the genetic improvement. Sometimes this is acceptable (e.g. a genetic defect is a valid reason for culling). In other cases such criteria are doubtful (e.g. “body volume” as an indicator of productivity) or not recommendable (e.g. frame size or “dairyness”).

3.3 Design of breeding scheme Designing a breeding programme requires taking a range of decisions in a logical order. The designer of the programme should be aware that such a process evolves over time – from the simple to increasing levels of sophistication as organization and capacity develop. Most of the decisions involve determining how best to utilize present population structure to reliably generate the improvement and/or restructuring that is needed. Economic evaluation is an integral part of this process, and should be carried out both for the pre-implementation phase and for evaluating the change being realized when the programme is underway. Investment decisions in the breeding programme should be assessed with respect to the three components contributing to the rate of genetic change: selection intensity, selection accuracy and generation interval. Based on these components, alternative scenarios are assessed. Theoretical knowledge of quantitative genetics is used to predict the gains to be expected from different scenarios (Falconer and Mackay, 1996). For this purpose, population genetic parameters

392

such as heritability and phenotypic variation of the traits are needed to build up the selection index (reasonable assumptions can also be made) (Jiang et al., 1999). A suitable mating plan is then outlined. It must allow sufficient records to be obtained for genetic evaluation, and sufficient elite animals to be produced for the nucleus and for multiplication in the lower levels of the breeding pyramid. Note that in performing these activities, the designer of the programme is already in the optimization phase. When designing the breeding programme, it should not be forgotten that most aspects are directly influenced by the reproductive rate of the breeding animals. A higher reproductive rate means that fewer breeding animals are needed. More offspring per breeding animal allows more accurate estimation of breeding value.

3.4 Data recording and management Recording of performance data and pedigrees is the main driving force for genetic improvement. Abundant and accurate measurements lead to efficient selection. In practice, however, resources are limited. The question then is: which traits should be measured and on which animals? Preferably, the traits included in the breeding objective should be measured, but this will depend on the ease and cost of measurement. The nucleus animals, at least, should be measured for performance and pedigree. The collection of performance data on which to base selection decisions is a vital component of any breeding programme, and it should be regarded as such, rather than as a by-product of recording systems primarily designed to assist short-term management (Bichard, 2002). The task of collecting, collating and using data in genetic evaluation requires good organization and considerable resources (Wickham, 2005; Olori et al., 2005). In many instances, special schemes may need to be put in place to generate and record the required data. The cost and complexity of these schemes vary depending on the type of breeding organization, the type of traits, and the method of testing.

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

Type of breeding organization. pig and poultry breeding companies have in-house facilities for the collection and storage of all required data, whereas other breeding organizations may rely on resources owned by more than one stakeholder. For example, this is the case in a typical dairy cattle breeding programme (see subchapter 4.1). Type of trait. When body weight of live animals is the trait of interest, all that is needed is a weighing scale. However, to measure feed efficiency in individual animals, more sophisticated equipment may be needed to allow the recording of individual feed intake. Performance versus progeny or sib testing. In a performance-testing scheme, the traits of interest are recorded directly in every individual. For example, body weight and growth are often recorded over a fixed period during the lifespan of beef cattle, pigs, broiler chickens or turkeys. Basically, a cohort of animals is managed together under similar conditions over a period of time during which individual performance is measured. This can be done on the farm, or at a performance test station where cattle or pigs from different herds or farms are brought together for a direct comparison under the same conditions. Sometimes, the information of interest may not be measurable directly in the selection candidate, either because the expression of the trait is sexlimited as in the case of milk and egg production, or because the traits can only be recorded after the death of the animal (e.g. carcass composition). In these circumstances, indirect recording by progeny and/or sib testing is required. This is also useful for traits with low heritability, which may require several records to accurately evaluate an individual. Progeny testing refers to a scheme in which an individual is evaluated on the basis of performance records obtained from its progeny. It is mainly associated with males (Willis, 1991), as it is easier to generate large numbers of progeny from a single male than from a single female. Typically, not all males are progeny tested, but only the males born from “elite matings”. Progeny testing is very useful to increase selection

accuracy for species with low reproductive rates, and to test genotype–environment interactions. For many ruminant species, the cost of a central progeny testing facility may be prohibitive. It is, therefore, a common practice to involve as many farmers or commercial producers as possible. The farmers are encouraged to accept semen from a group of young sires to be used on a proportion of their female animals. Because the young sires are not of proven genetic merit, farmers involved in progeny testing often require good incentives to participate (Olori et al., 2005). In these circumstances, the total costs (several hundred thousand US Dollars) are often borne by the owners of the young sire under test. Pedigree information. In addition to performance records, genetic evaluation in a breeding programme requires pedigree information. The quality of pedigree information depends on its depth and completeness. Whether the breeding objective involves genetic improvement or the prevention of extinction resulting from a loss of genetic variation, the pedigree of all breeding animals must be recorded and maintained. Information systems. When the resources are available, a centralized database with shared access has been shown to be beneficial and cost effective (Wickham, 2005; Olori et al., 2005). The provision of comprehensive management-related information from such a system often serves as a stimulus for further participation in data recording schemes. The requirement for small breeding programmes may simply be a single personal computer with adequate spreadsheet, data management and reporting software, while national-level programmes may require a specialized department utilizing modern information technology (Grogan, 2005; Olori et al., 2005).

3.5 Genetic evaluation Progress in a breeding programme requires that animals of superior genotypes for the traits of interest are identified and selected to breed

393

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

the next generation. Identifying these animals requires disentangling the environmental contribution from the phenotypic observation. This is accomplished by breeding value prediction or genetic evaluation. This is a core activity in every breeding programme. The genetic evaluation should be reliable. BLUP methodology, applied to a variety of models depending on the traits and data available, has become the standard method for nearly all species. The evaluation should also be available in time to make the best use of the investment in data collection and database management. A genetic evaluation system using BLUP relies on good data measurement and structure. If these prerequisites are in place, investment in BLUP is usually highly cost effective. Across-herd evaluation has the advantage of allowing fair comparisons of predicted breeding values (PBVs) of animals in different herds, which leads to selection of more animals from the genetically superior herds. To do this, genetic links (usage of animals across herds and across years) are critical. In order to use the information from different herds, an adequate organizational structure is needed. This can be achieved through close collaboration between breeders, their associations, and universities or research centres. Unique identification for all animals that supply data for the breeding scheme is essential. The data analysts, with guidance and assistance from breed association personnel, assign animals to contemporary groups (groups of animals of about the same age that are raised together with the same treatment). This assignment may be critical for accurate genetic evaluation. The breeders submit data to the association, and after checking for obvious errors, the information is forwarded to the evaluation team for analysis. For ruminants, the evaluations are performed once or twice a year, but for pig and poultry meat programmes, where the selection is performed on a monthly, weekly or bi-weekly basis, evaluations are run continuously. The results of the genetic predictions (PBV and aggregate indices) are typically printed on the

394

animals’ registration certificates. It is common to print PBVs in sale and semen catalogues. This means that the end users (farmers) have to understand and accept the EBVs that are produced, and know how to use them. There is no sense in running a genetic evaluation if the results are left untouched by the end users. A typical genetic evaluation unit requires both qualified staff, and adequate material resources to carry out data analysis and produce suitable reports to facilitate selection decisions. Many large-scale breeding programmes have a dedicated genetic evaluation unit in-house. However, it is also easy to contract this evaluation out to an external institution. Many universities and research centres provide a genetic evaluation service for national and non-national breeding programmes. Such services can cover several different breeds or species, as the principle of genetic evaluation and the software involved will be similar in each case. Perhaps, the most popular genetic evaluation unit with international repute is the International Bull Evaluation Service (INTERBULL). The centre, which is based at the Swedish Agricultural University in Uppsala, was set up as a permanent subcommittee of the International Committee for Animal Recording (ICAR), and provides international genetic evaluation to facilitate the comparison and selection of dairy bulls on an international scale. Another example is BREEDPLAN, a commercial beef cattle genetic evaluation service with an operational base in Australia, which has clients in many countries.

3.6 Selection and mating Selection should predominantly be based on the selection criterion. From each sex, as few breeding animals as possible should be selected to maximize selection intensity, with the only restrictions being the number of animals required for a minimum population size, and the number needed for reproductive purposes. As reproductive rates of males are generally much higher than those of females, far fewer breeding males than females are normally selected.

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

Selection candidates may be of different ages, and thus unequal amounts of information may be available about them. For example, older males may have a progeny test, while for younger ones, their own performance, or that of their dam or sibs, will be the only information available. If BLUP is used, such candidates can be easily and fairly compared. Selecting more animals with accurate EBVs, and only the very best animals with less accurate EBVs, is probably the best approach. It is widely accepted that the use of family information, as occurs in BLUP, increases the probability of co-selection of close relatives, which in turn leads to increased inbreeding. Various methods are used to reduce inbreeding while maintaining high rates of genetic gain. All these methods are based on the same principle – reducing the average relationship between the individuals selected. Computer programmes have been developed to optimize selection decisions for a given list of candidates for which pedigree information and EBVs are available. Ad hoc methods to control inbreeding include selecting a sufficient number of males, as the rate of inbreeding depends on effective population size; not overusing the males within the nucleus; restricting the number of close relatives selected, especially the number of males selected per family; limiting the number of females mated to each male; and avoiding mating between full and half sibs. These simple rules have been quite effective in maintaining a low level of inbreeding in commercial poultry and pig breeding. Mating of selected animals may or may not be at random. In the latter case, the very best of the selected males are mated to the very best of the selected females – this is known as assortative mating. The average genetic value of the progeny born in the next generation does not change, but there will be more variance among the progeny. When multiple traits are included in the breeding objective, assortative mating may be useful – matching qualities in different parents for different traits. Any mating strategy will require sufficient facilities. For natural mating, animals to be mated

have to be put together in the same paddock, but separated from other animals of reproductive age. AI can be used, but also requires a range of resources and expertise (semen collection, freezing and/or storing, and insemination).

3.7 Progress monitoring This involves the periodic evaluation of the programme with respect to progress towards the desired goal. If necessary, it leads to a reassessment of the goal and/or the breeding strategy. Monitoring is also important to ensure early detection of undesired effects of the selection process, such as increased susceptibility to diseases or a reduction in genetic variation. To assess progress, phenotypic and genetic trends are usually obtained by regressing average annual phenotypic and breeding values on year of birth. In addition to this information, breeders run regular internal and external performance testing. An external testing scheme needs to cover a wide range of production environments to ensure that selected animals can perform well under a wide range of conditions. Other sources of information, and probably the most important, are field results and feedback from customers. Ultimately, the customer is the best judge of the work done.

3.8 Dissemination of genetic progress The value of superior individuals is limited if they do not efficiently contribute to the improvement of the gene pool of the whole target population. The wide impact of genetic improvement depends on the dissemination of genetic material. Reproductive technologies, especially AI, are very important in this respect. However, their impact varies between species. In sheep and goat breeding, the exchange of genetic material largely depends on trade in live animals. In the case of cattle, AI allows bulls selected in the nucleus to be used across the whole population. In principle, there is no problem in allowing an exceptional bull to have many progeny throughout the population. However, performing AI using semen

395

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

from bulls from the same family very intensively will ultimately lead to inbreeding. It should be possible to apply the elements described above even under basic conditions. Breeding structures do not necessarily require sophisticated systems of data recording and genetic evaluation, nor do they initially require use of reproductive technologies. The breeding structure should be determined in accordance with what is possible and what is optimum. Environmental or infrastructure restrictions, traditions and socioeconomic conditions have to be considered when planning breeding programmes.

4

Breeding programmes in highinput systems

In high-input systems, continuous genetic improvement is generated mainly by straightbreeding within a breed or line. In the case of ruminants, this is largely a result of the strong position and active work of breeding associations, and of the spectacular results obtained by this method. Cross-breeding is used to realize the benefits of hybrid vigour (heterosis) and complementarity. In poultry and pigs, breeders concentrate their efforts on within-breed or line selection, and use cross-breeding to capitalize on heterosis for fitness traits and on complementarity for other traits. The number of livestock breeding companies in the world is relatively low, but they are of great economic significance. They increasingly operate on a global scale. As the following subchapters will illustrate, the structure, including the ownership, of breeding organizations differs greatly between species.

4.1 Dairy and beef cattle breeding Selection criteria In dairy cattle, the average milk, fat and protein production per cow per year has increased enormously in the past decades as a result of the

396

widespread use of breeds such as the HolsteinFriesian and intensive within-breed selection. This increase is also a reflection of the fact that productivity has for many years been an important selection objective, with selection mainly being based on production and morphological traits. Recent years have seen a growing concern on the part of consumers about animal welfare issues, and about the use of antibiotics in livestock production. Breeding organizations have also realized that selecting solely for product output per animal leads to a deterioration of animals’ health and reproductive performance, increased metabolic stress and reduced longevity (Rauw et al., 1998). As a result, emphasis on functional traits has increased, and less attention is paid to product output. Selection for functional traits is now based on direct recording of these traits rather than through type traits. Breeding values for a wide range of functional traits have been developed and applied in most countries. This enables breeding organizations and farmers to pay direct attention to these traits in their selection decisions.

Box 81 Calving problems in Belgian White Blue cattle In beef cattle, the demand for high-quality meat has led to the use of breeds, such as the Belgian White Blue, that have extreme phenotypes. However, this breed has an extremely high rate of caesarean sections (Lips et al., 2001). In the short term, this rate cannot be significantly reduced. The extreme muscularity of the Belgian White Blue is mainly caused by the myostatin gene, a single autosomal recessive gene which is located on chromosome 2. It is, therefore, questionable whether a reduction in calving difficulties can be realized while maintaining the extreme muscularity. Because of this, as well as the obvious animal welfare concerns, the future of the breed is questionable.

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

Box 82 Cross-breeding to address inbreeding-related problems in Holstein cattle

The Holstein breed, which is composed almost completely of American Holstein genes, has largely replaced other breeds of dairy cattle throughout much of the world. Production and conformation traits have been emphasized in the breeding of Holsteins because of moderately high heritability and ease of data collection. However, female fertility, calving ease, calf mortality, health and survival have been ignored until very recently. Problems related to functional traits, coupled with increased inbreeding on an international

scale, have resulted in tremendous interest in crossbreeding among commercial dairy producers. Purebred sires will continue to be sought to breed almost all dairy heifers and cows for cross-breeding. Most cross-breeding systems with dairy cattle will make use of three breeds to optimize the average level of heterosis across generations. For further information see: Hansen (2006).

TABLE 99 Breeding objectives in ruminants Objectives/product

Criteria

Further specification

Quantity

Milk carrier production

Contents/quality

Fat percentage, protein percentage, somatic cell count, milk coagulation

Growth rate

At different ages

Carcass quality

Fat content, bone/meat ratio

Meat quality

Tenderness, juiciness

Quantity Fibre quality

Length, diameter

Genetic defects

BLAD, mule foot and CVM

Production traits Milk

Beef

Wool

Functional traits Health and welfare

Mastitis incidence Udder conformation

Udder attachment, udder depth and teat traits

Feet and leg problems Locomotion

Indicator of hoof disorders

Reproduction efficiency

Female fertility Male fertility Calving ease Number of live offspring

Showing heat, pregnancy rate Non-return rate Direct and maternal effects, still births

Feed Efficiency

Feed conversion efficiency Milk production persistency

Workability

Milkability Behaviour

Longevity

Functional herd life

Milking speed

397

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

Breeders face difficulties in two areas – breeding (including recording) and marketing. With regard to breeding, there are problems associated with correlated responses to selection. In most cattle breeding programmes, an aggregate index is constructed that includes traits such as growth, milk yield, fertility, conformation, number of somatic cells in the milk, calving ease and duration of productive lifespan (for more details see Table 99). In dairy cattle, the main focus has been (and is still) put on milk yield, despite the negative genetic correlations between milk yield and reproduction and health-related traits. Undesired side-effects have, therefore, been observed – including lower fertility, and greater susceptibility to mastitis, leg problems and ketosis. In beef cattle and in sheep, selection for growth has led to higher birth weights and increasing risk of birth problems. Higher growth rates can also be expected to increase the mature size of breeding females. This may result in lower reproductive rates if larger animals are unable to meet their nutritional requirements because of limitations in the quantity or quality of the available forage. These undesired effects can be avoided, or at least reduced, by increasing the weight of functional traits within selection indices. This supposes that these traits can be directly measured. Recording of functional traits often remains an important bottleneck hindering their inclusion in breeding schemes. This is illustrated by the example of efficiency of feed utilization. Recording feed intake in a large number of animals is currently impossible – preventing efficient selection for this trait. There are also problems related to marketing. For milk, good management practices have been in place in many countries for a long time, and product quality has a direct impact on the price paid to producers. In the case of meat, however, traceability and organization in the production chain has traditionally been poor. This limits opportunities to improve quality. In general, farmers are not rewarded for meat quality, and often only poorly rewarded for carcass quality.

398

Organization and evolution of the breeding sector Because of the low reproductive rate, the long generation interval and the large amount of space required to house each animal, cattle breeding has a more complex and more open organizational structure than poultry or pig breeding. Gene flow can occur both from the breeder to the producer and vice versa. Information resources are shared between players at different levels. In a typical dairy cattle breeding programme, pedigree information is often recorded, owned and managed by breed societies, while milk production records are owned by farmers, but collected and managed by milk recording organizations. Information on fertility and reproductive performance are kept by companies that provide AI services, while health information generally resides with veterinarians. Often, these organizations are in decentralized locations and may store information in different systems. Because cattle production is a major traditional agricultural enterprise and because breeding has a major impact on this enterprise, cattle breeding programmes have more input from government agencies than do poultry or pig breeding, and therefore have a country-specific outlook. Most programmes were either initiated or sustained with support or grants from national government agencies (Wickham, 2005). Organizations such as the Animal Improvement Programs Laboratory (AIPL) of the United States Department of Agriculture (USDA), Canadian Dairy Network (CDN), Cr-Delta in the Netherlands, and l’Institut de l’Elevage (IE) in France, play major roles in cattle breeding programmes in their respective countries, especially in data management and genetic evaluation. This is also the case for breed societies, which have played a major role in maintaining and enhancing the integrity of their respective breeds. The success of the HolsteinFriesian, which is by far the dominant sire breed in most dairy herds in the Western world, is testimony to the activities of the World HolsteinFriesian Federation (WHFF). The formation of

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

Box 83 Norwegian Red Cattle – selection for functional traits The Norwegian Red (NRF) is a high-producing dairy cattle breed in which fertility and health have been included in a selection index (known as the Total Merit Index) which has been in operation since the 1970s. The case of the NRF provides a practical illustration that production and functional traits can be successfully balanced in a sustainable breeding programme. This achievement has been based on an effective recording system and a willingness to place sufficient weight on the functional traits. The programme is run by GENO, a cooperative owned and managed by Norwegian dairy farmers. Currently, ten traits are included in the Total Merit Index. The following list shows the relative weight given to each: Milk index 0.24 Mastitis resistance 0.22 Fertility 0.15 Udder 0.15 Beef (growth rate) 0.09 Legs 0.06 Temperament 0.04 Other diseases 0.03 Stillbirths 0.01 Calving ease 0.01 Key features of the programme include the fact that more than 95 percent of herds participate in the recording system and are on a computerized mating plan, 90 percent of matings are carried out using AI, and there is 40 percent use of test bulls. All diagnosis and health registration is carried out by veterinarians, and databases are maintained for pedigree and AI-related information. About 120 young bulls are tested annually with progeny groups of 250 to 300 daughters – thus enabling the inclusion of traits with low heritability (such as mastitis with a heritability of 0.03 and other diseases with 0.01) while still providing a selection index with high accuracy. Milk production per lactation in the best herds exceeds 10 000 kg, with the top cows producing more than 16 000 kg. The genetic trend is positive with

respect to fertility – the average 60 day non-return rate in the population is 73.4 percent. Between 1999 and 2005 incidence of mastitis in NRF cows was reduced from 28 percent to 21 percent, and it is estimated that of this reduction 0.35 percent per year was the result of genetic improvement. Major calving difficulties are reported in less than 2 percent of calvings, and less than 3 percent of calves are stillborn. The sustainability of the breeding programme is promoted by a number of factors: • Both production and function are expressed by many traits, and they are both strongly weighted in the breeding strategy. • Many different combinations can result in a high total breeding value. This allows for the selection of animals from different breeding lines and, thus, automatically reduces the risk of inbreeding. • The breeding work is based on data from ordinary dairy herds, which guarantees that the breeding programme produces animals that are well adapted to normal production conditions. Provided by Erling Fimland. For further information see: http://www.geno.no/genonett/ presentasjonsdel/engelsk/default.asp?menyvalg_id=418

Photo credit: Erling Fimland

399

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

herd books with dedicated members and the importance of show ring performance (which are strictly within-breed affairs) have helped sustain pure-breed development and the maintenance of all major breeds of dairy and beef cattle. The selection programmes conducted by AI centres have developed from local to national schemes, and are increasingly operating internationally. The dissemination of genetic material from “superior” animals is now global. It is predicted that within the next ten to 15 years AI centres will become unified into a few worldwide breeding companies, such as now exist in the pig and poultry sectors. For example, in the early 1990s the “Genus” breeding programme was the major cattle programme in the United Kingdom. Over the years, Genus has merged with ABS genetics from the United States of America to form a global company, which now supplies bovine genetics from a variety of dairy and beef cattle breeds to over 70 countries. More recently, Genus bought Sygen, a biotech company. Breeding programmes in cattle rely on commercial producers to generate sufficient data for genetic evaluation. Data recording, therefore, takes place in all tiers of the breeding pyramid. This requirement is greatest in the case of dairy programmes, which require large progeny groups for the accurate evaluation of bulls (especially for traits with low heritability), or in beef cattle to be able to estimate direct and maternal effects. The use of AI to disseminate semen across many herds is prevalent, and this helps to facilitate the comparison of animals raised in different environments. AI also enables higher intensity in the selection of males. Successful selection within dairy cattle breeds is the result of well-organized programmes for the measurement of production, testing of young bulls and effective genetic evaluation. The high level of feeding in commercial dairy production allows a high proportion of a cow’s genetic potential to be expressed, which in turn allows selection to be particularly effective.

400

Cross-breeding studies with dairy cattle have consistently found significant levels of heterosis between dairy breeds for milk production, fertility and survival traits. However, successful long-term selection for high levels of milk production in the Holstein-Friesian has led to the widespread use of straight-bred animals of this breed. However, increasing pressure from commercial producers, who are suffering losses related to poor fertility and longevity, and the need for flexibility in product development is likely in the future to lead to increased development of hybrid cattle at the breeding programme level. Cross-breeding applied to beef cattle is often undertaken without a well-designed programme. In beef cattle, cross-breeding programmes are difficult to implement in herds that use fewer than four bulls. Even for larger operations, managing the herds separately, as is required in organized cross-breeding programmes, can be difficult (Gregory et al., 1999). In cattle, the introduction of AI has resulted in an enormous reduction of the number of sires and contributed to the exchange of genetic material between regions and countries. Through AI, bulls selected in the nucleus are used in the general population. As a result of the high reproductive rate of sires, the selection of bulls contributes 70 percent to total genetic change in dairy and beef cattle populations.

4.2 Sheep and goat breeding Selection criteria Sheep and goats are kept for meat, milk, and wool or fibre (see Table 99 for corresponding breeding goals). Sheep milk is an important product in Mediterranean countries. It is mainly transformed into a variety of cheeses (e.g. Roquefort, Fiore Sardo, Pecorino Romano and Feta). Milk production and quality are important breeding criteria. Milk sheep may also be bred for growth rate, reproductive traits such as twinning rate, and type traits such as udder shape (Mavrogenis, 2000). Conversely, in northwestern Europe, meat

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

is the most significant product obtained from sheep. Specific breeding objectives will depend on the production environment (e.g. mountain vs. lowland), and may include growth rates, carcass quality, reproductive performance and maternal abilities. Commercial wool production is dominated by Australia and New Zealand with their specialized flocks of straight-bred fine-wool sheep of the Merino type. Although the animals all descend from the Merino sheep of Spain, different strains have been developed over the years. The need for animals adapted to specific environmental conditions has shaped breed development. In Australia, for example, different strains of Merino have been bred for their adaptation to the environment in different parts of the country. With respect to wool production, criteria for selection normally include clean fleece weight and fibre diameter. Increasing economic importance of meat relative to wool has led to a shifting of breeding objectives towards criteria such as reproduction rate and sale weight. In Mediterranean countries, in South Asia, and in parts of Latin America and Africa, goats are mainly kept for their milk. In Mediterranean countries and in Latin America, goat milk is often used for cheese production, whereas in Africa and South Asia, it is consumed raw or acidified. In other parts of Asia and Africa, goats are kept mainly for meat production. In these regions very little supplemental feeding is provided, and browse provides a significant amount of the nutritional requirements. The animals are of moderate to small size, and of moderate to light muscling. An exception is the development of the Boer goat for meat production in South Africa. The breed has been introduced to other countries in Africa and to other parts of the world such as Australia.

Organization of the breeding sector Major breeding programmes for fine-wool sheep are based in the southern hemisphere (Australia and New Zealand). These programmes are based on straight-breeding. However, in fine-wool sheep

operations where a significant part of the income is from lambs (for slaughter), self-contained F1 production has been used. Under this type of programme, all ewes are straight-bred for fine wool. A large fraction of the selected ewes are mated to fine-wool rams to produce replacement females. The remaining ewes are mated to terminal sires and all the lambs are sold. In the case of meat sheep breeding, the average size of flocks is generally too small to allow intensive within-flock selection. This problem has been overcome through cooperative breeding schemes. Nucleus breeding schemes are well established (e.g. James, 1977), but sire-referencing schemes (SRS) have recently gained popularity. In SRS, genetic links are created between flocks by mutual use of specific rams (reference sires). These connections allow comparable acrossflock genetic evaluation, offering a larger pool of candidates for selection for collective goals. About two-thirds of performance-recorded sheep in the United Kingdom, including all of the major specialized meat breeds, now belong to these schemes (Lewis and Simm, 2002). Cross-breeding is the basis of the stratified sheep industry of the United Kingdom (Simm, 1998). The system functions on the basis of a loose structure involving several breed societies, government agencies and other institutions. Traditional hill breeds such as the Scottish Blackface are straightbred under the harsh production conditions of the hills. Ewes from these pure breeds are sold to farmers in “upland” areas (where the climate is less harsh and there is better grazing). Here, they are crossed with rams from intermediate crossing breeds such the Blueface Leicester. F1 Females are sold for breeding in lowland flocks where they are mated to terminal-sire breeds such as the Suffolk and the Texel. Most data recording and genetic evaluation aim at improving the terminal-sire breeds to produce rams of superior genetic quality. Data recording and genetic evaluations are carried out by commercial operations such as Signet or by research institutions supported by public funds.

401

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

Most dairy goats are in developing countries. However, breeding programmes are concentrated mainly in Europe and North America. The French selection programme, based on AI with frozen semen and oestrus synchronization (60 000 goats inseminated/year), and the Norwegian programme, based on rotation of sires in several herds (buck circles), are examples of organized progeny testing programmes. They include a formal definition of selection objectives and organized mating to produce young sires and their progeny. Probably, the best example of a structured meat goat breeding programme is that run by the Boer Goat Breeders’ Association of Australia. Cashmere and mohair production is based on straight-breeding of the respective breeds. There is almost no cross-breeding involving Angoras.

4.3 Pig and poultry breeding Selection criteria in pigs As in the case of ruminants, pig breeding programmes have been very successful in achieving genetic improvement of economically important traits, especially daily gain, backfat thickness, feed efficiency and, during the last decade, litter size (for more details see Table 100). At present, the goal is to breed for more robust and efficient animals to meet different environmental conditions. This implies finding an adequate strategy to deal with genotype × environment interaction, and the placing of more emphasis on secondary traits which have up to the present been of negligible economic importance. Secondary traits include piglet survival, interval between weaning and first oestrus, longevity of sows, conformation (especially legs), vitality of pigs until slaughter weight, meat colour and

TABLE 100 Breeding objectives in pigs Objectives

Criteria

Further specification

Growth rate

At different ages

Production traits

Carcass weight Carcass quality

Uniformity, leanness of carcass

Meat quality

Water holding capacity, colour, flavour

General resistance

Robustness

Vital piglets Survival of pigs

Maternal ability, teat number

Stress

Elimination of stress (halothane) gene in dam lines, and where possible, in male lines

Congenital effects

Examples: atresia ani, cryptorchism, splay leg, hermaphrodism and hernia

Leg problems

Leg weakness and lameness.

Litter size

Number of slaughter pigs per sow per year

Functional traits Health and welfare

Efficiency

Feed conversion efficiency Longevity

402

Functional herd life

Lifetime production with minimal health problems

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

TABLE 101 Breeding objectives in poultry Objectives/product

Criteria

Further specification

Egg number

Number of saleable eggs per hen

External egg quality

Average egg weight, shell strength and colour

Internal egg quality

Egg composition (yolk/albumen ratio), firmness of albumen and freedom from inclusions (blood and meat spots)

Growth rate

Weight gain; age at market weight

Carcass quality

“Yield” in terms of valuable parts, especially breast meat; select against breast blisters and other defects to reduce condemnation rate

Disease resistance

Not routinely used

Production traits Egg

Meat

Functional traits Health and welfare

Monofactorial genetic defects Leg problems in broilers and turkeys Osteoporosis in laying hens Heart and lung insufficiency

Incidence of “sudden death syndrome” and ascites in broilers and “round heart” in turkeys

Cannibalism, feather pecking

Feed efficiency

Feed consumption per: • kg egg mass in laying hens, • kg weight gain in broilers and turkeys Residual feed consumption

Longevity

Length of productive life

drip loss. The health of the pigs is becoming more important. This means not only improving the sanitary status in breeding farms, but also selecting for general disease resistance under commercial conditions. As in the case of ruminants, there are some difficulties involved in implementing efficient selection for “functional” traits. There are still no appropriate tools to select for better resistance to diseases or to reduce metabolic disorders. Sufficient knowledge of the genetic aspects of welfare is lacking. Stress recording methods need to be improved – for example, through the use of non-invasive methods for measuring

stress-indicating parameters, determination of catecholamine levels, and heart-rate recording on under-skin chips. Improved knowledge of the cognitive abilities and coping strategies of pigs might enable individual characteristics to become indicative of ability to adapt to various housing conditions and social challenges, and could be included in selection criteria. Additionally, there is a need for further assessment of the impact of selection for specific disease resistance and welfare objectives.

403

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

Selection criteria in poultry Laying hens have been selected mainly for productivity. Over several decades, breeding programmes were refined, and more and more traits were included in the selection objectives. Today, the main selection objectives are: the number of saleable eggs per hen housed per year, efficiency of converting feed into eggs, external and internal egg quality, and adaptability to different environments (for more details see Table 101). For poultry meat, substantial genetic improvements in terms of market weight at a younger age and correlated feed efficiency have been achieved by simple mass selection for juvenile growth rate and “conformation”. During the 1970s, direct selection for efficient feed conversion was introduced. During the last two decades, the emphasis of selection has shifted increasingly to traits that are of primary importance to processing plants – breast meat yield, total carcass value, efficiency of lean meat production, uniformity of product, and low mortality and condemnation rates. The development of specialized male and female lines, and the introduction of controlled feeding of parents, are effective tools to overcome the negative correlation between juvenile growth rate and reproductive traits. The most obvious challenges for the poultry industry are related to diseases. Primary breeding companies have eliminated egg-transmitted disease agents such as leucosis virus, mycoplasms and Salmonella from their elite stock, and continue to monitor freedom from these problems. Other diseases such as Marek’s disease, E. coli, Campylobacter coli, and highly pathogenic avian influenza are more difficult to control. In the field of animal welfare, the main challenges for breeders are to adapt laying hens to alternative management systems – for example, to reduce feather pecking and cannibalism in non-cage systems (pecking and cannibalism are also serious problems for turkeys and waterfowl), and to reduce the incidence of cardio-vascular insufficiencies (sudden death syndrome and

404

ascites) and leg problems in broilers and turkeys. However, the causes of these problems are probably multifactorial, and further research is required.

Organization and evolution of pig and poultry breeding sectors The modern poultry industry has a typical hierarchical structure with several distinct tiers. Breeding companies based mainly in Europe and North America, with subsidiaries in major production regions, own the pure lines. They have to keep the whole production chain in mind – hatcheries, egg and meat poultry growers, processing plants, retailers and consumers. Hatcheries (multipliers) are located near population centres around the world. They receive either parents or grandparents from the breeders as day-old chicks, and produce the final crosses for egg producers and broiler, turkey or duck growers. Today, egg processing plants, slaughterhouses and feed suppliers have developed contractual relationships with egg producers and poultry growers, which provide the latter with better financial security, but at the cost of reduced initiative and freedom. The pig sector has a similar pyramidal structure, which is largely the result of the introduction of cross-breeding, AI and specialized breeding farms. However, some differences exist between the pig and the poultry sectors. For example, a pig producer will typically obtain the “commercial” animals by mating sows from a specialized dam line and boars from a specialized sire line – both genders being bought from the breeding company (and not from a multiplier as in poultry). In contrast to poultry, there are still breeding associations for pigs, and national genetic evaluation is performed. While genetic evaluations for the large breeding companies may be performed in-house, genetic evaluations at the pure-breed level are conducted by governmental institutions (e.g. by the National Swine Registry in the United States of America) or breed associations.

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

Pig and poultry breeding schemes are sometimes referred to as “commercial” breeding programmes because of the corporate ownership structure of these companies. Over the years, these programmes have amalgamated to become large corporations. In poultry, for example, only two to three groups of primary breeders account for about 90 percent of the layers, broilers and turkeys produced annually. Furthermore, some of these companies are owned by the same group. The pig breeding industry has more breeding companies and fewer large ones (such as PIC and Monsanto), but is following the same trend. The recent entry of the giant Monsanto into this sector is a clear indication of this tendency. Because of the competitive nature of the business and the high level of investment, “commercial” breeding companies are usually at the forefront in the application of technologies. These leading companies are on the verge of incorporating genomic information in their breeding programmes, at a time when many breeders are merely discussing the feasibility of the approach. The activities of these commercial breeding companies are characterized by the following features: • Pedigree selection occurs in the nucleus only. • Selection is strictly within specialized lines (or breeds). These lines are designated as sire and dam lines and are selected with different intensities. In poultry bred for meat and in pigs, male lines are selected for growth and lean meat production, while female lines are selected for reproduction. New lines are constantly developed either by crossing between existing lines or by further selection in a given direction. • The final product is a cross between two or more pure-bred lines. For economic reasons, each breeding company will sell under several trademarks (accumulated through acquisitions and fusions), but will in fact only have a limited number of differentiated products. Indeed, pig or poultry breeding companies develop lines to meet few (two or

three) breeding goals, which vary depending on the extent of their global market share and the degree of variation in the production environments in which the clients operate. For example, a breeder may develop a high-yielding, fast-growing line for use under high-input conditions where superior-quality feed allows the expression of the animals’ full genetic potential, and a line for more challenging environments that is more “robust”, but has lower performance for production traits.

5

Breeding programmes in lowinput systems

5.1 Description of low-input systems Many of the world’s livestock will continue to be kept by smallholders and pastoralists. These producers often have limited access to external inputs and to commodity markets. Even if external inputs are locally available, there is usually little cash available for their purchase. To quote LPPS and Köhler-Rollefson (2005): “Cash products are often of secondary importance, especially in marginal and remote areas. Traditional breeds generate an array of benefits that are more difficult to grasp and to quantify than outputs of meat, milk, eggs or wool. These include their contribution to social cohesion and identity, their fulfilment of ritual and religious needs, their role in nutrient recycling and as providers of energy, and their capacity to serve as savings bank and insurance against droughts and other natural calamities.” The livestock owned by smallholders and pastoralists may be autochthonous or originate from early introductions of exotic breeds to the area. Traditional livestock keepers have no technical training in genetics and many are illiterate. However, they possess valuable local knowledge about breeds and their management. They have breeding goals and strategies even if they are not “formalized” or written down. For example, they may share breeding males (they

405

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

Box 84 Community-based sheep management in the Peruvian Andes

Agriculture in the central Andes of Peru is severely limited by low temperatures and drought, and most rural households depend on livestock for their income. Rangeland sheep are economically the most important species, and are used as a source of food, as a means of obtaining goods through exchange, and to generate cash through the sale of live animals or wool. To a lesser extent they are also used for cultural activities, recreation and tourism. Criollo sheep represent 60 percent of the Peruvian sheep population. They are mainly raised on family farms and by individual farmers, who value the local breed highly. A dual-purpose breed, developed from a cross between Criollo sheep and Corriedale sheep imported from Argentina, Australia, Chile, New Zealand and Uruguay between 1935 and 1954, is also available. Peasant farmers maintain both the Criollo and the composite breed.

In this part of Peru, peasant communities have organized themselves independently to improve the management of their sheep, with little support from the government. Multicommunal and communal enterprises, cooperatives, as well as family and individual farms, are common. Farmers exchange genetic material, experiences and technologies. Multicommunal and communal enterprises have far higher production rates than individual farmers. They have successfully set up participatory breed improvement programmes based on open-nucleus schemes, are technically efficient, keep their pastures in good condition, and use some of their profits to improve the social well-being of their members – for example, by buying school materials, selling milk and meat at reduced prices, and providing assistance to the elderly. Provided by Kim-Anh Tempelman. For further information see: FAO (2007).

seldom have more than one of a given species) with their neighbours or the entire community. In conclusion, formalizing genetic improvement in these conditions is a challenging, but definitely not an impossible or inappropriate, task.

5.2 Breeding strategies It is important to keep in mind that whatever strategy is considered, it will be successful only if certain conditions are met. Meeting these conditions does not guarantee success, but neglecting them will certainly lead to failure. The owners of the livestock should be involved as much as possible, and preferably from the very beginning of the programme. The social structure of the region and the objectives of the producers should be carefully taken into consideration. The whole system, and not only one element of it, needs to be considered. For example, when considering a cross-breeding scheme in a remote

406

area, it is necessary to ensure that the progeny of cross-bred animals are viable in these conditions. The programme should be as simple as possible. In some cases it may be feasible to cross-breed individual females to males from other breeds that are available in the vicinity, but programmes that require continuous use of males of more than one breed are not feasible under low-input systems.

Breeding strategies Determining the breeding objectives is the most important and difficult task in any genetic improvement programme, and there is even less margin for error in low-input systems. The questions that need to be considered under these conditions include: what (if anything) should be changed, and what would actually be an improvement in these conditions? A low-input system is also a low-output system, but this does not necessarily mean low productivity.

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

Box 85 Genetic improvement of an indigenous livestock breed – Boran cattle in Kenya The Boran, a medium-sized cattle breed of East African origin, is the breed most widely kept primarily for beef production in the semi-arid zones of Kenya. Commercial ranchers prefer the Boran to Bos taurus breeds because of their relative adaptability to the local environment – achieved through generations of natural and artificial selection in conditions of high ambient temperature, poor feed quality, and high disease and parasite challenge. Boran genetic material is recommended as a means of improving beef production in other indigenous and exotic breeds in the tropics. Genetic exports to Zambia, the United Republic of Tanzania, Uganda, Australia and the United States of America occurred from the 1970s to the 1990s. Export of Boran embryos to Zimbabwe and South Africa took place during 1994 and 2000. This market potential has been an incentive for farmers to improve the breed. By the 1970s, the Boran had undergone cross-breeding with B. taurus types, backcrossing, and within-breed selection (which was mainly based on visual appraisal guided by experience). During the 1970s a recording scheme was initiated. Producers sent animal performance records routinely to the Livestock Recording Centre (LRC) for genetic evaluation. However, because of inconsistency and delays in the release of evaluation results, and the expenses associated with recording, most producers opted out of the scheme. In 1998, a bull performance testing project was implemented by the National Beef Research Centre in an attempt to evaluate bulls across various herds. However, the performance testing could not be sustained because of a lack of funds. Recently, breeding objectives for Boran production systems have been developed. Systems are classified according to the sale age of the animals (24 or 36 months), levels of input (low, medium or high), and final goal (beef or dual purpose). Traits of economic importance have been identified, and genetic parameters have been estimated for some of them.

These traits include sale weight for steers and heifers, dressing percentage, consumable meat percentage, milk yield in dual purpose production systems, cow weight, cow weaning rate, cow survival rate, postweaning survival rate, and feed intake of steers, heifers and cows. Genetic improvement of the Boran in Kenya is facilitated by the Boran Cattle Breeders’ Society (BCBS). Membership of the society is restricted to farmers keeping Boran cattle, and other interested stakeholders. At present, the activities of the society focus on administration, maintaining breed standards, and searching for new markets for both beef and genetic material. Farmers are still independent with respect to selection and genetic improvement. Occasional exchange of genetic material between herds as a means of preventing inbreeding is probably the only form of interaction between farms. On most farms, selection focuses largely on weaning weights and calving interval. To evaluate their animals, some farmers have purchased various computer programmes to enable them to re-orientate on-farm performance recording to suit their management purposes. The BCBS is among the most active breeders’ associations in Kenya. It is not at present subsidized financially, but is involved in strategic cooperation with the LRC which stores and evaluates performance records for those producers still participating in the recording scheme. The BCBS also cooperates with the National Agricultural Research System in the exchange of information – especially on nutrition and breeding. Research aimed at developing appropriate genetic improvement programmes for the Boran and updating the current ones is ongoing. Provided by Alexander Kahi. For more information on Boran cattle and BCBS see: www.borankenya.org

407

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

Box 86 A llama breeding programme in Ayopaya, Bolivia

In the high Andes of Bolivia, llama keeping is an important and integral part of the mixed farming practised by rural households. Llamas provide smallholders with dung, meat and fibre; they are used as pack animals and also play an important social role. Llamas, as an autochthonous species, contribute to maintaining the ecological balance of the fragile local ecosystem. There are two main types of llama – the “Kh’ara” type, and the wool type known as “Th’ampulli”. The region of Ayopaya (department of Cochabamba) where the breeding programme takes place is situated at 4 000 to 5 000 metres above sea level in the eastern Cordillera of the Andes. Because of the geographical conditions and very basic infrastructure, the region is difficult to access. In 1998, a breeding programme for llamas was jointly initiated by the 120-member local producers association ORPACA (Organización de Productores Agropecuarios de Calientes), the NGO ASAR (Asociación de Servicios Rurales y Artesanales) and two universities (University Mayor de San Simon, Cochabamba, and University of Hohenheim, Germany). Initial funding was assured by the above-mentioned institutions. Continuation of the programme critically depends on securing external funding. Llamas in Ayopaya region

Photo credit: Michaela Nürnberg

408

Restraining llamas for transport

Photo credit: Michaela Nürnberg

As a first step, the production system was studied by participative observation and the use of questionnaires. The phenotype of 2 183 llamas of the Th’ampulli type was also characterized. The process revealed that the llamas possess fibre of extraordinarily high quality – 91.7 percent fine fibres and a fibre diameter averaging 21.08 μm. This fibre quality is unmatched by other llama populations in Bolivia. The animals, therefore, constitute a unique genetic resource. Interviews with representatives of the textile industry and traders provided information on the economic potential of the fleece. The performance of identified llamas was recorded and breeding parameters estimated. A mating centre run by ASAR to which members of ORPACA bring their females for service was established in Calientes in 1999. Selected males are kept at the centre during the mating season. The phenotypic evaluation of the males aims to identify animals with uniform fleece colour; a straight back, legs and neck; testicles that are of equal size and not too small; and no congenital defects. Six communities within a radius of about 15 km are served by the mating centre. Performance data for the offspring are recorded by trained farmers. • continues

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

Box 86 cont. A llama breeding programme in Ayopaya, Bolivia

Functions of llamas and breeding objectives are being recorded, ranked and valued jointly with the llama keepers. In a stepwise procedure, the breeding programme is being adapted to meet the breeders’ preferences, the market conditions, and the biological constraints. Genetic progress has not yet been evaluated because of the llama’s long generation interval.

Linear measurements on llamas

Provided by: Angelika Stemmer, André Markemann, Marianna Siegmund-Schultze, Anne Valle Zárate. Further information can be obtained from the following sources: Alandia (2003); Delgado Santivañez (2003); Markemann (forthcoming): Nürnberg (2005); Wurzinger (2005), or from: Prof. Dr Anne Valle Zárate, Institute of Animal Production in the Tropics and Subtropics, University of Hohenheim, 70593 Stuttgart, Germany. E-mail: [email protected]

Photo credit: Javier Delgado

Llama herd (of Emeterio Campos) in Ayopaya region

Deworming during sire selection at Milluni

Photo credit: André Markemann

Photo credit: André Markemann

409

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

Box 87 Pastoralists’ breeding criteria – insights from a community member Criteria for breeding decisions (in order of importance) A breeding bull should: • be active and agile – so as to serve all the females in the herd in a given breeding period (it is considered that such bulls are tolerant of diseases and parasites, and that diseases in them are easily detected);

4

The East African pastoralists of the Karamoja cluster keep a range of livestock including Zebu cattle, Small East African goats, Persian Black Head sheep, grey donkeys and light brown dromedaries. Some also keep indigenous chickens. Uses of livestock are diverse, and include food; a store of wealth, and a currency against which other commodities can be valued; a source of recreation and prestige; a means for the payment of debts, fines and compensations; a means of transport and agricultural traction; a source of skins and fibres; and a source of dung for fuel, fertilizer or building. Livestock also have many cultural roles such as being given to the bride’s family at the time of marriage. They are also slaughtered at the time of rituals associated with births; funerals; the onset of transhumance; rain-making; averting bad omens, epidemics or enemy attack; cleansing ceremonies; or curing an ailment on the prescription of a village herbalist. Criteria for breeding decisions are multifaceted, and reflect the interaction of social, economic and ecological factors. They include not only productivity, but also the taste of meat, blood, and milk; agreeable temperament; coat colour; religious requirements; disease and parasite resistance; mothering instincts; walking ability; tolerance of droughts; survival on meagre feed; and tolerance of extremes of temperature or precipitation.

• produce offspring that can maintain their body weight (and milk yield in the case of females) even during periods of feed shortage; • have large body size and weight – important for marketability and status, but be not too heavy to perform its breeding functions; • be tall, with a wide chest and straight back – again to meet breeding functions; • have the coat colour or horn configuration identified with the owner5 or the community; • have a coat colour and quality suitable for marketing or other uses; • have good temperament – aggressive6 towards predators, but not towards other livestock or humans; • bulls kept to breed offspring for draught purposes should have large body weight, and be strong and tractable; • breeding bulls should stay in the owner’s herd, graze well, and not be fond of roaming or fighting other bulls. • continues

4“

Karamoja Cluster”: The entire Ateker people in Uganda, Kenya, Ethiopia and the Sudan who generally share a common livelihood. “Ateker” people: (variously called “Ngitunga/Itunga” = the people). The people with a common origin living in Uganda (NgiKarimojong including Pokot, Iteso), Kenya (NgiTurukana; Itesio, Pokot); Ethiopia (NgiNyangatom/NgiDongiro) and in the Sudan (NgiToposa) and their neighbours; who speak similar languages and refer to their clans as Ateker (pl. Ngatekerin/ Atekerin). Some clans of Ateker people are spread all over Karamoja cluster.

410

5

Pastoralists also base their own name on the colour or horn configuration of their favorite bulls. This is typical in the Karamoja Cluster. Such names have the prefix Apa- which means “the owner of the bull with a ... coat colour/horn configuration”. For instance, the name “ApaLongor” means “the man with a bull with a brownish coat colour”. The favourite breeding bull receives many privileges from the owner such as being adorned with a bell, or prompt treatment when ill. 6

Indiscriminate aggression is unacceptable in livestock, even if other traits are favourable.

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

Box 87 cont. Pastoralists’ breeding criteria – insights from a community member Female breeding animals should: • have a stable high milk yield that is not only tasty and has ample butterfat content, but is also able to maintain healthy and quick growth of the offspring; • be able to calve regularly and produce quickgrowing offspring; • be tolerant of disease, heat, cold and long droughts; • survive on little feed and maintain high milk yield, particularly in the dry season when the feed quantity and quality is low; • the udder should be wide and the teats always complete;

The world should appreciate the role pastoralists play in sustainably utilizing their uniquely adapted breeds. Not only do these animals provide food and income security for their keepers, but they also contribute to the maintenance of genetic diversity, thereby providing a resource for future genetic improvement programmes. In this regard, pastoralists need appropriate support from livestock services provided by national governments, civil society organizations and the international community. Provided by Thomas Loquang (member of the Karimojong pastoralist community). For further information see: Loquang (2003); Loquang (2006a); Loquang (2006b); Loquang and Köhler-Rollefson (2005).

• cows should be docile to humans and other livestock, but aggressive towards predators; • small stock (goats, sheep) should regularly give birth to twins7.

For the low-input system, it is inadequate to think of genetic improvement only in terms of increases in output traits, such as body weight, milk or egg production, or fleece weight. Efficiency is also a key criterion. Unfortunately, very little is known about the genetic improvement of intrinsic efficiency. Increased efficiency is usually measured in terms of increased gross efficiency. The increased gross efficiency observed in highproducing animals results from the fact that a lower proportion of the animals’ nutrient intake is used for maintenance, and a correspondingly 7

Please note that it is a taboo for small ruminants to deliver twins at the first delivery. It is allowed only in the subsequent births. Similarly, it is a taboo for cattle to deliver twins whether at the first or subsequent delivery. Any such situations (births of twins) would lead to the animals concerned being slaughtered by stoning or beating. An animal in this situation is said to have become a witch and as such should be promptly eliminated!

higher proportion is used for production. This does not mean that the animal needs less feed to achieve a given level of performance. Selection based on residual feed intake (RFI) has been proposed as a means of improving intrinsic efficiency. This is an important criterion for all species and all production systems. Genetic selection to reduce RFI can result in animals that eat less without sacrificing growth or production performance (Herd et al., 1997; Richardson et al., 1998). For example, in contrast to the ratio of weight gain/feed intake, residual feed consumption is relatively independent of growth. RFI is therefore a more sensitive and precise measurement of feed utilization (Sainz and Paulino, 2004).

411

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

Box 88 The Bororo Zebu of the WoDaaBe in Niger – selection for reliability in an extreme environment This example refers to cattle breeding in a specialized pastoral system in Niger. The WoDaaBe are full-time cattle keepers. Marketing livestock is the cornerstone of their livelihood strategy. Their herds contribute a substantial proportion of national cattle exports, particularly to the large markets of Nigeria where Bororo animals sell at a premium. “Extreme environment” here refers to a combination of a harsh ecosystem characterized by stochastic events, and comparatively poor access to both primary resources and external inputs. WoDaaBe herders exploit a semi-arid territory characterized by erratic and unpredictable rainfall. In an ordinary year, fresh grass is available for only two to three months at any given location. Access to forage, water and services requires a degree of purchasing power and negotiation with neighbouring economic actors competing for these resources. The WoDaaBe are usually on the weaker side in these transactions. It has been proposed that the concept of “reliability” is key to understanding the management strategies of pastoralists under such conditions (Roe et al., 1998). “High-reliability” pastoral systems are geared to the active management of hazards rather than their avoidance, with the aim of ensuring a steady flow of livestock production. In these systems, breeding has to be closely interconnected with the environment and the production strategy. The main goal of the WoDaaBe is to maximize the health and reproductive capacity of the herd throughout the year. Their management system aims to ensure that the animals eat the highest possible amount of the richest possible diet all year round (FAO, 2003). This involves specialized labour, focusing on managing the diversity and variability of both grazing resources and livestock capabilities.

Photos credit: Saverio Krätli

• continues

412

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

Box 88 cont. The Bororo Zebu of the WoDaaBe in Niger – selection for reliability in an extreme environment The nutritional value of the range is maximized by moving the herd across zones that show spatially and temporally heterogeneous distribution of fodder. Additionally, the animals’ capacity as feeders is stretched beyond the natural level. While feeding capacity has in part a genetic base (for example the enzymatic system or the size and conformation of the mouth), it can also be greatly affected by learning, based on individual experience and imitation between social partners (for example efficient trekking and grazing behaviour and diet preferences). Animals’ feeding motivation is manipulated through optimizing their digestive feedback, and ensuring best fodder quality and preferred foraging conditions. A carefully diversified diet of grasses and browse is favoured, in order to correct nutritional imbalances which, particularly during the dry season, could keep feeding motivation low by triggering negative digestive feedback. The dry-season watering regime is also tailored in order to hone cattle’s digestive performance to meet the herders’ long-term strategic goal of maximizing reproduction. The production strategy is very demanding on both people and the herd. With the onset of the dry season, while other pastoral groups sharing the same ecosystem move closer to water points, where water is more accessible but pasture is poor, the WoDaaBe move in the opposite direction, trying to keep their camps close to prime fodder. This results in longdistance mobility and a watering regime which, at the peak of the hot season, often involves journeys of 25–30 kilometres to reach the well, with the herd drinking every third day. It is, therefore, essential to the WoDaaBe’s production strategy that functional behavioural patterns are maintained within the herd. Consequently, their breeding system focuses on fostering social organization and interaction within the herd. It encourages sharing of animals’ feeding competence across the breeding network, and tries to guarantee the genetic and “cultural” continuity of successful cattle lineages within the network. These

lineages have proved capable of prospering under the WoDaaBe’s herd management system, and over a long enough period to have included episodes of severe stress. The breeding strategy focuses on ensuring the reliability of the herd’s reproductive performance, more than on maximizing individual performance in specific traits. Breeding involves selective mating of cows with matched sires, and a marketing policy that targets unproductive cows. Less than 2 percent of the males are used for reproduction. Close monitoring of the herd allows early detection of oestrus and ensures that more than 95 percent of births result from matchmaking with selected males. A different sire is used for almost every oestrus of a particular cow, with an overall ratio of about one sire every four births. Pedigree sires are borrowed across large networks of (often related) breeders. Sire borrowing remains frequent (affecting about half the births) even when a breeder owns pedigree sires of his own. Matchmaking with non-pedigree sires, owned or borrowed, affects about 12 percent of births. Both practices are maintained explicitly in order to preserve variability. Matrilineal genealogies and the sire of each animal in the herd are usually remembered, together with pedigrees of special sires, and the identity and owner of all borrowed sires. A cow’s productivity depends heavily on how well the animal responds to the management system. By adopting a production strategy that manipulates the animals’ experience of the ecosystem, the herder exposes his animals to diverse natural environments involving particular combinations of favourable and unfavourable foraging and watering conditions. Over the years, some cows prosper and produce a numerous progeny while others die or struggle and are sold. In this way, the WoDaaBe are able to harness natural selection pressure for their breeding purposes. Provided by Saverio Krätli. For more information see: Krätli (2007).

413

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

Box 89 Community-driven breeding programmes for local pig breeds in north Viet Nam

In the mountainous areas of Northwest Viet Nam, livestock breeding and management programmes, can contribute to improving rural livelihoods if they respect the production objectives, intensity and resource-availability of the area’s resource-poor smallholder mixed farming systems. The local Ban pig which shows considerable hardiness, but has a low reproductive and growth performance is increasingly being replaced by higher-yielding Vietnamese Mong Cai sows from the Red River Delta. In a collaborative project between the National Institute of Animal Husbandry (NIAH) Hanoi and the University of Hohenheim, Germany8, communitybased pig breeding programmes have been established in seven villages, differing in terms of their remoteness and market access. A total of 176 households currently participate in the programmes. On-farm performance testing schemes have been developed. Farmers are provided with data sheets on which they record the performance of their pigs (mainly date of farrowing and number of piglets). Vietnamese and German researchers cross-check data and collect additional data by weighing and identifying animals when they visit the villages. Specially trained farmers enter the data into the project databank using the PigChamp® software and researchers analyse the data. Farmers in Viet Nam often receive money for their participation in projects; in the case of this project, compensations are gradually being reduced. Results are fed back to farmers at seminars/training modules, and are further used to optimize breeding (gilt selection and optimization of mating plans). In order to ensure long-term sustainability, local partners such as the province Department of Agriculture and Rural Development (DARD) and the sub-Department of Animal Health of Son La province, are actively involved and trained. Cooperation with provincial 8

Funded by the German Research Association (DFG) in the frame of the Thai-Vietnamese-German collaborative research programme SFB 564 and by the Ministry of Science and Technology, Viet Nam.

414

extension services will be strengthened in the current project phase. In earlier phases, the service’s strong orientation towards intensive management in favoured regions meant that exchanges were limited. Financial support for the future of the project seems to be available thanks to NIAH’s official mandate to carry out projects on AnGR conservation. Moreover, the marketing element of the current project is aimed at ensuring long-term economic viability. Initial performance testing results indicate that Mong Cai and their cross-bred offspring (sired by exotic boars) are more suited to semi-intensive, marketoriented production conditions, where the higher levels of inputs needed to achieve higher production can be provided. They seem to be less robust in the harsh upland climates and under conditions of low and varying input intensity. Ban pigs are only suited for the extensive conditions of subsistence-oriented resourcepoor farming. As the project continues, efforts are being made to further develop breeding goals, to optimize stratified breeding programmes, and to implement marketing programmes. Close to town, lean meat is produced from the cross-bred offspring of Mong Cai sows. Production of Ban pigs continues in remote locations with pure or cross-bred animals marketed as a branded speciality – contributing to the “conservation through use” of this local breed.

Provided by Ute Lemke and Anne Valle Zárate. Further information can be obtained from the following sources: Huyen, et al. (2005); Lemke, (2006); Rößler. (2005), or from: Prof Dr Anne Valle Zárate, Institute of Animal Production in the Tropics and Subtropics, University of Hohenheim, 70593 Stuttgart, Germany. E-mail: [email protected]

• continues

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

Box 89 cont. Community-driven breeding programmes for local pig breeds in north Viet Nam BẢN ÐỒ HÀNH CHÍNH GIAO THÔNG TỈNH SƠN LA

Mong Cai sow

Ban fatteners

Photos provided by Ute Lemke

Compensation Farmers Data sheets

Researchers (German, Vietnamese) Additional data recording Information

Feedback of results Training, advice Controlled mating

Optimize pig management

Centrally managed project data bank Software: PigChamp ® Data entry: farmers (Vietpignew software) Management: researchers

Breeding management, controlled mating

Answer research question

• continues

415

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

Box 89 cont. Community-driven breeding programmes for local pig breeds in north Viet Nam

Pigs in Song Ma district

Weighing pigs in Pa Dong, Mai Son district

Photo credit: Pham Thi Thanh Hoa

Photo credit: Regina Rößler

Data recording in low-input systems

Breeding schemes

The absence of a credible recording scheme and resources for adequate data storage and management hinder the development of sustainable breeding programmes in low-input systems. Running a computerized database can be expensive and may require specialized skills. The absence of technical skills and financial resources has been identified as the main obstacle to the establishment of sustainable animal recording systems in many African countries (Djemali, 2005). Continuous advances in information technology mean that data recording devices are becoming cheaper and offer greater potential for recording in low-input systems. The use of hand-held devices, laptops and the Internet could make it easier for small numbers of people to gather and transmit large amounts of data from remote locations to a central database. Such a database could be based in a university or a government department. Provision of facilities of this type is one way in which governments or donor agencies could facilitate the development of breeding programmes for low-input systems in developing countries.

If genetic change is justified, how can it be achieved? The choice is between straight or crossbreeding, but choosing the appropriate option is far from simple. In low-input systems, adaptation to the environment is a prerequisite for improved efficiency. This is a matter of great importance, as intervention to reduce environmental stresses (supplementary feeding, parasite treatments or other management inputs) is often unaffordable. In these circumstances, straight-breeding to improve well-adapted indigenous breeds may be an option. Implementing a straight-breeding programme is a long-term undertaking, requiring considerable resources, good organization, and (most of all) commitment of all stakeholders. These requirements tend to be lacking under low-input systems in the developing world, and programmes that do exist are only of a very limited scope. For example, most controlled breeding of the West African Dwarf Goat has been in research institutions (especially in those in Nigeria) (Odubote, 1992). Cross-breeding with an exotic breed may appear to be a more rapid means to improve performance

416

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

with a minimal increase in inputs. However, the higher performance of the cross-breeds is accompanied by higher nutritional and management requirements (disease control, housing, etc.). Therefore, any system that incorporates higher-performing crossbred animals will require (among other needs) more feed resources – which in many cases can only be achieved by maintaining a smaller number of animals. If, after careful analysis, cross-breeding is considered to be a better option than straightbreeding the local breed, the programme should be developed in a way that can be sustained with locally available inputs. Cross-breeding with an exotic (non-adapted) breed presents particular difficulties. Even if the F1 animals are sufficiently adapted, the pure-bred exotic males will usually be under environmental stress, and this will

Box 90 The cost of heterosis Heterosis has sometimes been referred to as a free opportunity for increased profitability. Although it may be worth more than it costs, heterosis is not free. It involves at least two types of costs. First, there is the cost involved in meeting the nutritional requirement for the additional performance. The higher performance of the crossbred animal tends to reduce the cost per unit of production, because the cost for maintenance becomes a smaller fraction of the total requirement, but there is a cost for the extra production. A second type of cost is associated with potential changes in population structure. These costs may include (1) reductions in the size (and a corresponding increase in the level of inbreeding) of an original pure-bred population which occurs because of the need to accommodate the cross-bred population, and (2) a reduced opportunity to select for female productivity in a population where some of the cross-bred females are not considered to be candidates for selection (as in any terminal-sire system).

often result in a reduced reproductive life. Even if the male of the exotic breed can be successfully maintained, the backcross resulting from mating F1 females with the exotic males will almost always lack adequate adaptation to the area. Therefore, the F1 females should preferably be mated to adapted-breed sires. One option under these conditions is to use F1 males, generation after generation. Under such a system, the original local females are mated to F1 males, resulting in offspring that are 1/4 exotic. These quarter-blood females are, in turn, mated to F1 males, resulting in females that are 3/8 exotic. After a few generations the animals would be very close to half exotic. This system introduces exotic influence into the population, but never uses or produces any animals that are more than half exotic. Another option for cross-breeding under lowinput systems is to cross different breeds that are well adapted to the production conditions. The obvious advantage of such programmes is the ability to maintain and produce the breeding stock in the area without additional inputs. It

Box 91 Nigeria’s Village Poultry Improvement Scheme A Village Poultry Improvement Scheme aimed at upgrading the indigenous breed of chicken with improved exotic breeds (Rhode Island Red, Light Sussex and Australorp) was initiated in Nigeria around 1950 (Anwo, 1989). The strategy was to cull all indigenous males and replace them with improved imported breeds in a “cockerel exchange programme” (Bessei, 1987). This scheme failed because the cross-bred chicks, though better in performance, could not survive in the semi-wild extensive backyard production system under which the indigenous chickens were raised. Another major drawback was that breed replacement resulted in a rapid loss in genetic variation and narrowing of the available AnGR.

417

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

Box 92 A community-based and participatory dairy goat cross-breeding programme in a low-input smallholder system in the eastern highlands of Kenya FARM Africa’s Meru project in Kenya provides an example of a comprehensive and flexible crossbreeding programme. Improved goat genotypes accompanied by improved husbandry practices have been adopted by very poor farmers with incomes well below US$1 per person per day. The local goats (Galla and East African) were proving difficult to maintain on small and declining farm sizes (0.25 to 1.5 acres), and the farmers had started to abandon goat production. Consequently, the cross-breeding programme aimed to provide more docile and productive animals. Sixtyeight female and 62 male British Toggenburg goats were imported from the United Kingdom and crossed with indigenous goats: the Toggenburgs providing the dairy potential and the local goats providing adaptability. Previous introductions and trials had indicated that Toggenburgs were better adapted than other exotic dairy breeds such as Saanens or AngloNubians. The project adopted a group and communitybased approach. The farmers established the project’s rules, by-laws and mechanisms. It was linked to the government, NARS, and international research institutes, which provided training in husbandry (housing, nutrition, fodder production, record keeping and healthcare), group dynamics, marketing and entrepreneurship.

Farmer groups initially comprised 20 to 25 members, but some lost members over time while others grew. Four such groups were linked in a unit (mainly for administrative and monitoring purposes), with representatives being elected to a larger body the Meru Goat Breeders’ Association (MGBA). Small (one buck and four does) breeder units were provided (as a loan to be paid back in kind) to one group member, who produced the Toggenburgs (T) needed for breeding stock. One pure-bred Toggenburg buck was provided to each farmer group and kept in a buck station, maintained by another group member. Local does were brought to the buck station for service. The resulting F1 female cross-breeds were backcrossed to unrelated Toggenburg bucks to produce ¾ Toggenburg and ¼ Local (L) animals. These were evaluated, and superior males selected to start new buck stations, where they were used to serve unrelated females of similar genetic composition (¾ T and ¼ L). Initial trials had shown that such does produced adequate amounts of both milk and meat, and were reasonably adapted to the local conditions. Through the MGBA, which also registered the cross-breeds with the Kenya Stud Book, groups rotated the bucks every 1 to 1.5 years to avoid inbreeding. Farmers who wished to further upgrade towards the Toggenburg had the

Project statistics 1996 to 2004 1996

1997

1998

1999

2000

2001

2002

2003

2004

New farmer groups

10

34

20

6

12

10

7

18

8

New buck stations

10

34

10

11

6

16

14

3

22

New breeders units

5

Buck services Families participating Cross-breeds produced

250

20

25

10

12

6

2

4

7

809

1 994

3 376

3 936

3 892

3 253

5 660

6 500

1 100

1 125

1 400

1 550

1 700

2 050

2 050

2 650

990

2 894

3 241

3 817

3 736

4 187

5 865

7 200

Source: FARM-Africa Dairy Goat and Animal Healthcare Project; six-monthly reports January 1996–June 2004.

• continues

418

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

Box 92 cont. A community-based and participatory dairy goat cross-breeding programme in a low-input smallholder system in the eastern highlands of Kenya opportunity to do so by further backcrossing the ¾ T females to unrelated pure T bucks. Two years after FARM Africa’s pull-out the number of operating groups has continued to increase. In 2006 the MGBA has 3 450 members, all of whom keep improved goats which produce between 1.5 and 3.5 litres of milk per day. The group produces about 3 500 litres of milk daily, some of which is processed and packaged for sale. Member families own more than 35 000 improved goats of which 30 percent have reliable pedigree and performance records. The performance records are used for calculating growth rates and milk yields. These data were formerly processed by FARM-Africa. After the phasing out of the project, MGBA has been encouraged to establish collaboration with universities and research institutions to support them in data processing. Most of the owners of the improved goats are no longer “poor”. Some have used profits from goat production to purchase one or two dairy cows, build better houses and educate their children. Production of yoghurt and fresh pasteurized milk (adding value) is indicative of scope for further developments..

The features that made the scheme successful include: • a farmer-based approach since its inception; • an emphasis on capacity building so that farmers can manage the programme; • availability of locally produced breeding material; • a group approach – farmers train each other and share experiences; • capacity building for extension staff, farmercentred extension messages, and participatory approaches; and • the community-based establishment of breeder units and buck stations. The scheme has ensured that after the end of “the project”, farmers are not reliant on government services. Breeding stock is supplied by farmers themselves, and a parallel animal healthcare service has also been established by training communitybased animal health workers, with links to more qualified paraveterinarians and veterinarians. An integrated fodder and reforestation programme was also established. Provided by Okeyo Mwai and Camillus O. Ahuya. For further reading see: Ahuya et al. (2004); Ahuya et al. (2005); Okeyo (1997).

would be logical to assume that such crosses would produce less-productive animals and/or exhibit less heterosis than crosses between a local and an exotic breed. However, Gregory et al. (1985) report estimates of heterosis for weight of calf weaned per cow of 24 percent between Boran and Ankole cattle, and 25 percent between Boran and Small East African Zebu.

With any cross-breeding scheme it is important to consider the whole system and all outputs produced. Commenting the value of the European dairy x Zebu F1 cow for milk production in the tropics, LPPS and Köhler-Rollefson (2005) write “in India, many owners of cross-bred cows cannot see a use for male calves, so let them die.”

419

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

6

Breeding in the context of conservation

Conservation programmes for AnGR are discussed in greater detail elsewhere in this report. The following discussion, therefore, focuses on aspects of breeding that need to be considered when implementing conservation measures. A conservation programme may simply aim at ensuring the survival of a population through monitoring and maintaining its integrity, or a programme may also have the objective of improving the performance of the population.

6.1 Methods for monitoring small populations FAO has produced several publications on the management of at-risk small populations – see for example FAO (1998). These documents provide a more extensive review of the subject. Where the Where the objective is merely to ensure the survival of the population and the maintenance of its integrity (as a pure population), the conservation strategy is limited to monitoring the population, and ensuring that inbreeding and effective population size are within acceptable limits. Inbreeding is the result of mating related animals. In a small population, all animals in future generations will come to be related to each other, and mating among these animals will result in inbreeding. The genetic effect of inbreeding is increased homozygosity – the animal receives the same alleles from both its parents. The degree of inbreeding and homozygosity in future generations can be predicted from the population size. As there is almost always a much smaller number of breeding males than breeding females, the number of breeding males is the more important factor determining the amount of inbreeding. The effective population size (Ne) is a function of the number of breeding males and breeding females. If Nm represents the number of breeding males and Nf represents the number of breeding females, effective population size can be calculated as: Ne = (4NmNf) / (Nm + Nf)

420

If the number of breeding males is the same as the number of breeding females, the effective population size is the same as the actual population size; if the numbers of males and females are different, the effective population size is less than the actual population size. If the number of breeding females is much larger than the number of males, the effective population size will be slightly less than four times the number of males. A decrease in effective population size in livestock populations can be observed in two situations. The first and most obvious case is when the actual population size decreases. This can result from the replacement of a significant proportion of a breed with breeding animals of another breed, or from cross-breeding a significant fraction of the breed. The second situation is when a particularly popular sire and his sons and other descendants are heavily used. From the time of the first establishment of breed societies up to the mid1900s, much of the popularity of particular sires came about as a result of success in the show ring. In more recent times, predicted genetic value for particular traits has been the decisive factor. In dairy cattle, selection was for many years almost entirely focused on milk yield. Hansen (2001) reports that although over 300 000 head were registered by the Holstein Association USA Inc. in 2000, the effective population size was only 37 head. Using pedigree records of cattle born in 2001, Cleveland et al. (2005) report an estimated effective population size in the American Hereford of 85 head. The American Hereford Association registered over 75 000 head in 2001. The level of inbreeding in a given population is dependent on effective population size rather than actual population size. The increase in the level of inbreeding per generation is expected to be 1/2Ne. This is the increase expected per generation if each animal produces an equal number of offspring, and the animals in the initial population are not related to each other. If these assumptions are not met, the degree of inbreeding will be higher. Based on this

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

relationship, Gregory et al. (1999) recommend that at least 20 to 25 sires be used per generation. This would also be a reasonable number to be used in the conservation of a breed. The use of 25 sires per generation would result in a rate of increase in inbreeding of about 0.5 percent per generation. While the loss of effective population size is an important issue in the conservation of AnGR, it is interesting to note that successful breeders have always accepted some level of inbreeding in their programmes. These breeders established herds or flocks that met their standards – the animals produced in these closed herds or flocks inevitably came to be closely related, and inbreeding resulted (Hazelton, 1939).

6.2 Conservation through breeding The objectives of a conservation programme may include not only ensuring the survival and integrity of the target population, but also improving its reproductive rate and performance while maintaining its specific adaptive features. Much of the above discussion of breeding strategy for low-input systems is likely to be applicable in these circumstances. This subchapter focuses on the potential risks associated with cross-breeding in the context of breed conservation. One option to safeguard a breed is to use it as one of the components of a cross-breeding programme. However, any use of pure-bred females to produce cross-breeds will reduce the population size unless there is a reproductive surplus of females. In many cases, the environmental and management conditions do not allow for much reproductive surplus – especially in cattle, which have low reproductive rates. As such, most of the females that are raised must be retained as breeding animals in order to maintain the size of the population. In fact, the largest effect comes from the requirement for a smaller number of indigenous breeding males, brought about by the smaller number of indigenous females that are being used to produce pure-bred offspring. A logical starting point for consideration of a crossbreeding programme is, therefore, to estimate

the amount of reproductive surplus in females. This can be measured in terms of the fraction of young females that are available for slaughter or for sale out of the programme (or region). As an example, for fairly well-managed beef herds in temperate areas, about 40 percent of the heifer calves are needed for replacements in order to maintain the size of the herd. With knowledge of the reproductive surplus of females, and knowledge of the fraction of the total population that is currently made up of cross-breeds, the fraction of pure-breeds that can be utilized to produce F1s without further decreasing the population size of the pure breed can be calculated. As an example, if there is a 20 percent reproductive surplus of females and the current population is composed of 50 percent pure-breeds and 50 percent cross-breeds (includes any pure-bred females that are currently being used for cross-breeding), the population could move towards a composition of slightly more than 50 percent pure-breeds producing purebreeds, slightly more than 20 percent pure-breeds producing F1s, and slightly less than 30 percent F1 females, without any further reduction in the size of the pure-bred population that is producing pure-breeds. These values assume that none of the females produced by the F1 females are retained as breeding females; in reality, this would probably never occur.

7

Conclusions

Breeding methods and organization vary greatly between industrialized commercial production systems and subsistence-oriented low external input systems. The current organization of the breeding sector is a result of a long evolutionary process. The latest development is the spread of the industrialized breeding model, characteristic of the poultry sector, to other species. The industrialized breeding model uses state of the art techniques for genetic improvement. Breeding programmes are based mainly on straight-breeding and vary according to the

421

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

characteristics of the species. Breeding companies market their animals worldwide. This tendency, which is well established among “commercial” pig and poultry breeders, is increasingly the case for beef and dairy cattle. To select for robust animals that are able to cope with different environments, breeders run selection programmes across different environments and management systems. However, it is not possible to have animals that produce well everywhere and under all conditions. As such, different breeds or lines may be developed to meet demands in high-input systems. To date, little is known about the genetic aspects of adaptation. Scientists and breeding companies are expected to explore these matters further in their research and their breeding programmes in the coming years. In low external input production systems, animals kept by smallholders represent an important element of household food security and of the social fabric of village communities. To a large extent, smallholders and pastoralists keep local breeds. Genetic improvement in these conditions is a challenging, but not impossible, task. Detailed guidelines for the design and execution of sustainable breed utilization and improvement programmes for low external input systems are being developed and validated. Straight-breeding to adjust a local breed to the changing needs of producers is the most viable option not only to keep it in production and hence safeguard it, but also to improve food security and alleviate poverty. Another option is to use it as a component of a well-planned cross-breeding programme. In conjunction with the introduction of a breeding programme, attention should be given to the improvement of management conditions and husbandry practices. A common tendency in research related to breeding programmes for all species is an increasing focus on functional traits – in response to the growing importance given to factors such as animal welfare, environmental protection, distinctive product qualities and human health. Examples of functional traits include robustness, disease resistance and behavioural traits, fertility, efficiency of feed utilization, calving ease and

422

milkability. Generally, considered as secondary traits in high-input systems, functional traits are of great importance in low-input systems. Recording of functional traits, however, still remains an important bottleneck which hinders their inclusion in breeding schemes. Information is lacking on the genetic basis of disease resistance, welfare, robustness and adaptation to different environments. Nevertheless, the dairy cattle and pig industries have started to use DNA typing of single genes and genomics (SNPs) to screen breeding animals. This will support the expected shift towards breeding for functional and lifetime productivity traits. Because of the tendency for reduced use of chemical medications in the developed world, animals are required to have better resistance, or at least tolerance, to particular diseases and parasites. However, for economic and animal welfare reasons, it is very difficult to select for such animals using classical quantitative genetic approaches. High expectations are therefore placed on genomics. Some applications are already in use to eliminate genetic disorders with Mendelian inheritance. In the case of the more complex resistance traits for which genetic markers have been identified, such as Marek’s disease in poultry and E. coli in pigs, few if any breeding companies have implemented DNAbased selection. Welfare has become an important element in consumers’ perception of product quality, especially in Europe. The main challenges for breeders are to select for better temperament, and reduce foot and leg problems and the incidence of cardio-vascular problems (in poultry kept for meat production). The causes of these problems are multifactorial. The increasing importance of functional traits will require inclusion of a wider range of criteria in breeding programmes. Some of these criteria may be best met by local breeds. Characterization (phenotypic and molecular) and assessment of these breeds for important traits may allow the detection of some that have unique features. Their further development through breeding programmes would ensure that they remain

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

available for future generations. Unfortunately, the reality is a continuous loss of breeds and lines. The developed world (where the majority of concerted genetic improvement efforts are occurring) contributes directly or indirectly to this loss by concentrating on a very small number of breeds. The deletion of genetic lines that accompanies the worldwide reduction in the number of breeding companies via buy-outs has also played a major role.

References Ahuya, C.O., Okeyo, A.M., Mosi, R.O. & Murithi, F.M. 2004. Growth, survival and milk breeds in the eastern slopes of Mount Kenya. In T. Smith, S.H. Godfrey, P.J. Buttery, & E. Owen, eds. The contribution of small ruminants in alleviating poverty: communicating messages from research. Proceedings of the third DFID Livestock Production Programme Link Project (R7798) workshop for small livestock keepers. Izaak Walton Inn, Embu, Kenya, 4–7 February 2003, pp. 40–47. Aylesford, Kent, UK. Natural Resources International Ltd. Ahuya, C.O., Okeyo, A.M., Mwangi, N. & Peacock, C. 2005. Developmental challenges and opportunities in the goat industry: the Kenyan experience. Small Ruminant Research, 60: 197–206. Alandia, E.R. 2003. Animal health management in a llama breeding project in Ayopaya, Bolivia: parasitological survey. Institute of Animal Production in the Tropics and Subtropics, University of Hohenheim, Stuttgart, Germany. (MSc thesis) Amer, P.R. 2006. Approaches to formulating breeding objectives. In Proceedings of the 8th World Congress on Genetics Applied to Livestock Production, August 13–18. 2006. Belo Horizonte, MG, Brazil. Andersson, L., Haley, C.S., Ellegren, H., Knott, S.A., Johansson, M., Andersson, K., AnderssonEklund, L., Edfors-Lilja, I., Fredholm, M., Hansson, I., Hakansson, J. & Lundstrom, K. 1994. Genetic mapping of quantitative trait loci for growth and fatness in pigs. Science, 263: 1771–1774. Anwo, A. 1989. Ministerial speech. In E.B. Sonaiya, ed. Rural Poultry in Africa: proceedings of an international workshop, pp 8–9. Ile-Ife, Nigeria. Thelia House Ltd. Bessei, W. 1987. International poultry development. In Proceedings, 3rd International DLG symposium on poultry production in hot climates, June 20–24 1987. Hamelin, Germany Bichard, M. 2002. Genetic improvement in dairy cattle – an outsider’s perspective. Livestock Production Science, 75: 1–10.

423

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

Bijma, P., Van Arendonk, J.A. & Woolliams, J.A. 2001. Predicting rates of inbreeding for livestock improvement schemes. Journal of Animal Science, 79: 840–853. Cleveland, M.A., Blackburn, H.D., Enns, R.M. & Garrick, D.J. 2005. Changes in inbreeding of U.S. Herefords during the twentieth century. Journal of Animal Science, 83: 992–1001. Cunningham, E.P., Dooley, J.J., Splan, R.K. & Bradley, D.G. 2001. Microsatellite diversity, pedigree relatedness and the contribution of founder lineages to thoroughbred horses. Animal Genetics, 32: 360–364. Dawson, M., Hoinville, L., Hosie, B.D. & Hunter, N. 1998. Guidance on the use of PrP genotyping as an aid to the control of clinical scrapie. Scrapie Information Group. Veterinary Record, 142: 623–625. Dekkers, J.C.M. & Hospital, F. 2002. The use of molecular genetics in the improvement of agricultural populations. Nature, 3: 22–32. Delgado Santivañez, J. 2003. Perspectivas de la producción de fibra de llama en Bolivia. Potencial y desarrollo de estrategias para mejorar la calidad de la fibra y su aptitud para la comercialización. Institute of Animal Production in the Tropics and Subtropics, University of Hohenheim, Cuvillier, Göttingen, Germany. (PhD thesis) Dickerson, G.E. 1969. Experimental approaches in utilizing breed resources. Animal Breeding Abstracts, 37: 191–202. Dickerson, G.E. 1972. Inbreeding and heterosis in animals. In Proceedings of Animal Breeding and Genetics Symposium in honor of Dr. J.L. Lush, pp. 54–77. Blacksburg, Virginia. ASAS, ADSA. Djemali, M. 2005. Animal recording for low to medium input production systems. In M. Guellouz, A. Dimitriadou & C. Mosconi, eds. Performance recording of animals, state of the art, 2004. EAAP Publication No. 113, pp. 41–47. Wageningen, the Netherlands. Wageningen Academic Publishers.

424

Ducrocq, V. & Quaas, R.L. 1988. Prediction of genetic response to truncation selection across generations. Journal of Dairy Science, 71: 2543–2553. Falconer, D.S. & Mackay, T.F.C. 1996. Introduction to quantitative genetics. 4th Edition. London. Longman. FAO. 1998. Secondary guidelines for the development of national farm animal genetic resources management plans: management of small populations at risk. Rome. FAO. 2003. Know to move, move to know. Ecological knowledge among the WoDaaBe of south eastern Niger, by N. Schareika. Rome. FAO. 2007. Management of sheep genetic resources in the central Andes of Peru, by E.R. Flores, J.A. Cruz & M. López. In K-A. Tempelman & R.A. Cardellino eds. People and animals. Traditional livestock keepers: guardians of domestic animal diversity, pp. 47–57. FAO Interdepartmental Working Group on Biological Diversity for Food and Agriculture. Rome. Fernando, R.L. & Grossman, M. 1989. Marker-assisted selection using best linear unbiased prediction. Genetics Selection and Evolution, 21: 467–477. Fuji, J., Otsu, K. & De Zozzato, F. 1991. Identification of a mutation in porcine syanodine receptor associated with malignant hyperthermia. Science, 253: 448–451. Gregory, K.E & Cundiff, L.V. 1980 Cross-breeding in beef cattle: evaluation of systems. Journal of Animal Science, 51: 1224–1242 Gregory, K.E., Trail, J.C.M., Marples, H.J.S. & Kakonge, J. 1985. Heterosis and breed effects on maternal and individual traits of Bos indicus breeds of cattle. Journal of Animal Science, 60: 1175–1180. Gregory, K.E., Cundiff, L.V. & Koch, R.M. 1999. Composite breeds to use heterosis and breed differences to improve efficiency of beef production. Technical Bulletin. No. 1875. Springfield, Virginia. USDA Agricultural Research Service, National Technical Information Service. Groen, A.F. 2000. Breeding goal definition. In S. Galal, J. Boyazoglu & K. Hammond, eds. Developing breeding strategies for lower input animal production environments. Rome. ICAR.

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

Grogan, A. 2005. Implementing a PDA based field recording system for beef cattle in Ireland. In M. Guellouz, A. Dimitriadou & C. Mosconi, eds. Performance recording of animals, state of the art, 2004. EAAP Publication No. 113, pp. 133–140. Wageningen, the Netherlands. Wageningen Academic Publishers. Hanotte, O., Ronin, Y., Agaba, M., Nilsson, P., Gelhaus, A., Horstmann, R., Sugimoto, Y., Kemp, S., Gibson, J., Korol, A., Soller, M. & Teale, A. 2003. Mapping of quantitative trait loci controlling trypanotolerance in a cross of tolerant West African N’Dama and susceptible East African Boran cattle. Proceedings of the National Academy of Science USA, 100(13): 7443–7448. Hansen, L.B. 2001. Dairy cattle contributions to the National Animal Germplasm Program. Journal of Dairy Science, 84(Suppl. 1): 13. Hansen, L.B. 2006. Monitoring the worldwide genetic supply for cattle with emphasis on managing crossbreeding and inbreeding. In Proceedings of the 8th World Congress on Genetics Applied to Livestock Production, August 13–18. 2006. Belo Horizonte, MG, Brazil. Hazelton, J. 1939. A history of linebred Anxiety 4th Herefords of straight Gudgell & Simpson breeding. Kansas City, MO. George W. Gates Printing Co. Herd, R.M., Arthur, P.F., Archer, J.A., Richardson, E.C., Wright, J.H., Dibley, K.C.P. & Burton, D.A. 1997. Performance of progeny of high vs. low net feed conversion efficiency cattle. In Proceedings of the 12th Conference of the Assocation for the Advancement of Animal Breeding and Genetics, Dubbo, Australia, pp. 742–745. Hill, W.G. 2000. Maintenance of quantitative genetic variation in animal breeding programmes. Livestock Production Science, 63: 99–109. Hunter, N. 1997. Molecular biology and genetics of scrapie in sheep. In L. Piper & A. Ruvinsky, eds. The genetics of sheep, pp. 225–240. Oxon, UK. CAB International,

Huyen, L.T.T., Rößler, R., Lemke, U. & Valle Zárate, A. 2005. Impact of the use of exotic compared to local pig breeds on socio-economic development and biodiversity in Vietnam. Stuttgart, Beuren, Germany. James, J.W. 1972. Optimum selection intensity in breeding programmes. Animal Production, 14: 1–9. James, J.W. 1977. Open nucleus breeding systems. Animal Production, 24: 287–305. Jiang, X, Groen, A.F. & Brascamp, E.W. 1999. Discounted expressions of traits in broiler breeding programs. Poultry Science, 78: 307–316. Kennedy, B.W., Quinton, M. & van Arendonk, J.A. 1992. Estimation of effects of single genes on quantitative traits. Journal of Animal Science, 70: 2000–2012. Krätli, S. 2007. Cows who choose domestication. Cattle breeding amongst the WoDaaBe of central Niger. Institute of Development Studies, University of Sussex, Brighton, UK. (PhD thesis) Lamb, C. 2001. Understanding the consumer. In Proceedings of the British Society of Animal Science, 2001, pp. 237–238. Lande, R. & Thompson, R. 1990 Efficiency of markerassisted selection in the improvement of quantitative traits. Genetics, 124: 743–756. Larzul, C., Manfkedi, E. & Elsen, J.M. 1997. Potential gain from including major gene information in breeding value estimation. Genetics Selection Evolution, 29: 161–184. Lemke, U. 2006. Characterisation of smallholder pig production systems in mountainous areas of North Vietnam. Institute of Animal Production in the Tropics and Subtropics, University of Hohenheim, Germany. (PhD thesis) Le Roy, P., Naveau, J., Elsen, J.M. & Sellier, P. 1990. Evidence for a new major gene influencing meat quality in pigs. Genetical Research, 55: 33–40. Lewis, R.M. & Simm, G. 2002. Small ruminant breeding programmes for meat: progress and prospects. In Proceedings of the Seventh World Congress on Genetics Applied to Livestock Production, held August 19–23, 2002, Montpellier, France.

425

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

Lips, D., De Tavernier, J., Decuypere, E. & van Outryve, J. 2001. Ethical objections to caesareans: implications on the future of the Belgian White Blue. In Proceedings of the Third Congress of the European Society for Agricultural and Food Ethics, Florence, Italy, October 3–5 2001, pp. 291–294. Loquang, T.M. 2003. The Karamojong. In I. KöhlerRollefson & J. Wanyama, eds. The Karen Commitment: Part 2. The role of livestock and breeding; community presentations. Proceedings of a Conference of Indigenous Communities on Animal Genetic Resources. League for Pastoral Peoples and Endogenous Development and Intermediate Technology Development Group Eastern-Africa, Karen, Nairobi, Kenya, 27–30 October 2003. Bonn, Germany. German Non-Governmental Organisations Forum on Environment and Development. Loquang, T.M. 2006a. Livestock Keepers’ Rights. Paper presented at the side event during the Fourth Ad Hoc Open-Ended Intercessional Working Group on Article 8(j) and Related Provisions of the Convention on Biological Diversity, COP 8, Granada, Spain, 23–27 January 2006. Loquang, T.M. 2006b. The role of pastoralists in the conservation and sustainable use of animal genetic resources. Paper presented at the International Conference on Livestock Biodiversity, Indigenous Knowledge and Intellectual Property Rights; League for Pastoral Peoples and Endogenous Development, Rockefeller Study and Conference Centre, Bellagio, Italy, 27 March–2 April 2006. Loquang, T.M. & Köhler-Rollefson, I. 2005. The potential benefits and challenges of agricultural animal biotechnology to pastoralists. Paper presented at the Fourth All Africa Conference on Animal Agriculture, Arusha, Tanzania, 19–26 September 2005. LPPS (Lokhit Pashu-Palak Sanstham) & KoehlerRollefson, I. 2005. Indigenous breeds, local communities: documenting animal breeds and breeding from a community perspective. Sadri, Rajasthan, India. Lokhit Pashu-Palak Sanstham.

426

Markemann, A. (forthcoming). Development of a selection programme in a llama population of Ayopaya region. Department Cochabamba, Bolivia, Institute of Animal Production in the Tropics and Subtropics, University of Hohenheim, Germany. (PhD thesis) Mavrogenis, A.P. 2000. Analysis of genetic improvement objectives for sheep in Cyprus. In D. Gabiña, ed. Analysis and definition of the objectives in genetic improvement programmes in sheep and goats. An economic approach to increase their profitability, pp. 33–36. Zaragoza, Spain. CIHEAM–IAMZ. Meuwissen, T.H.E. 1997. Maximizing response to selection with a predefined rate of inbreeding. Journal of Animal Science, 75: 934–940. Nürnberg, M. 2005. Evaluierung von produktionssystemen der Lamahaltung in bäuerlichen gemeinden der Hochanden Boliviens. Institute of Animal Production in the Tropics and Subtropics, University of Hohenheim, Cuvillier, Göttingen, Germany. (PhD thesis) Odubote, I.K. 1992. Genetic and non-genetic sources of variation in litter size, kidding interval and body weight at various ages in West African Dwarf Goats. Obafemi Awolowo University, Ile-Ife, Nigeria. (PhD thesis) Okeyo, A.M. l997. Challenges in goat improvement in developing rural economies of Eastern Africa, with special reference to Kenya. In C.O. Ahuya & H. van Houton, eds. Goat development in East Africa. Proceedings of a workshop held at Izaak Walton Inn, Embu, Kenya, 8–11 December l997, pp. 55–66. Nairobi. FARM-Africa. Olori, V.E., Cromie, A.R., Grogan, A. & Wickham, B. 2005. Practical aspects in setting up a National cattle breeding program for Ireland. Invited paper presented at the 2005 EAAP meeting in Uppsala, Sweden. Pharo, K. & Pharo, D. 2005. Direction vs. destination. Pharo Cattle Co. Spring 2005 Sale Catalog, pp. 72–73.Cheyenne Wells, Colorado, USA. Pharo Cattle Co.

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

Rauw, W.M., Kanis, E., Noordhuizen-Stassen, E.N. & Grommers, F.J. 1998. Undesirable side effects of selection for high production efficiency in farm animals: a review. Livestock Production Science, 56: 15–33. Richardson, E.C., Herd, R.M., Archer, J.A., Woodgate, R.T. & Arthur, P.F. 1998. Steers bred for improved net feed efficiency eat less for the same feedlot performance. Animal Production Australia, 22: 213–216. Rößler, R. 2005. Determining selection traits for local pig breeds in Northern Vietnam: smallholders’ breeding practices and trait preferences. Institute of Animal Production in the Tropics and Subtropics, University of Hohenheim, Germany. (MSc thesis) Rocha, J.L., Sanders, J.O., Cherbonnier, D.M., Lawlor, T.J. & Taylor, J.F. 1998. Blood groups and milk and type traits in dairy cattle: After forty years of research. Journal of Dairy Science, 81: 1663. Roe E., Huntsinger, L. & Labnow, K. 1998. High reliability pastoralism. Journal of Arid Environments, 39(1): 39–55. Sainz, R.D. & Paulino, P.V. 2004. Residual feed intake. Agriculture & Natural Resources Research & Extension Centers Papers, University of California. Simianer, H. 1994. Current and future developments in applications of animal models. In Proceedings of the 5th World Congress on Genetics Applied to Livestock Production. Guelph. Canada. Vol. 18, pp. 435–442.

van Arendonk, J.A.M. & Bijma, P. 2003. Factors affecting commercial application of embryo technologies in dairy cattle in Europe – a modelling approach. Theriogenology, 59: 635–649. Wickham, B.W. 2005. Establishing a shared cattle breeding database: Recent experience from Ireland. In M. Guellouz, A. Dimitriadou & C. Mosconi, eds. Performance recording of animals, State of the art, 2004. EAAP Publication No. 113, pp. 339–342. Wageningen, the Netherlands. Wageningen Academic Publishers. Willis, M.B. 1991. Dalton’s introduction to practical animal breeding. 3rd ed. Oxford, UK. Blackwell Science Ltd. Woolliams, J.W. & Bijma, P. 2000. Predicting rates of inbreeding in populations undergoing selection. Genetics, 154: 1851–1864. Woolliams, J.W., Bijma, P. & Villanueva, B. 1999. Expected genetic contributions and their impact on gene flow and genetic gain. Genetics, 153: 1009–1020. Wurzinger, M. 2005. Populationsgenetische analysen in Lamapopulationen zur implementierung von leistungsprüfung und selektion. University of Natural Resources and Applied Life Sciences (BOKU), Vienna. (PhD thesis)

Simm, G. 1998. Genetic improvement of cattle and sheep. Tonbridge, UK. Farming Press, Miller Freeman UK Limited. Smits, M.A., Barillet, F., Harders, F., Boscher, M.Y., Vellema, P., Aguerre, X., Hellinga, M., McLean, A.R., Baylis, M. & Elsen, J.M. 2000. Genetics of scrapie susceptibility and selection for resistance. In Proceedings of the 51st Meeting of the European Association for Animal Production (EAAP). 21–24 August, The Hague, Paper S.4.4. EAAP. Rome

427

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

Section E

Methods for economic valuation 1

Introduction

The large number of AnGR at risk in developing countries, together with the limited financial resources available for conservation and sustainable use, means that economic analysis can play an important role in ensuring an appropriate focus for conservation and genetic improvement efforts. In this regard, important tasks include, inter alia: • determining the economic contribution that AnGR make to various sectors of society; • supporting the assessment of priorities through the identification of cost-effective measures which might be taken to conserve livestock diversity; and • assisting in the design of economic incentives and institutional arrangements for the promotion of AnGR conservation by individual farmers or communities. Swanson (1997) notes that human societies have been expanding and developing over time through a process involving biodiversity depletion. This process can be understood in terms of a tradeoff between maintaining the stock of diverse biological resources, and the benefits to human society derived from the depletion of this stock. AnGR erosion can, thus, be seen in terms of the replacement of the existing slate of livestock with a small range of specialized “improved” breeds. Such replacement occurs not only through substitution, but also through cross-breeding and the elimination of livestock because of production system changes. Genotype choices and threats to AnGR, therefore, need to be understood in the

context of the evolution of production systems (including biophysical, socio-economic and markets changes). See Part 2 for a further discussion of trends in livestock production systems. From an economic point of view, AnGR erosion can be seen as a result of drivers generating a bias towards investment in specialized genotypes, which in turn results in under-investment in a more diverse set of breeds. Economic rationality suggests that investment decisions will be determined by the relative profitability of the two options (assuming risk neutrality and well-functioning markets). However, from a farmer’s perspective, the relevant rates of return are those that accrue to him/her rather than to society or the world as a whole. To the farmer, the loss of a local breed will appear to be economically rational in a situation where the returns from the activities that lead to the loss are higher than those from activities compatible with genetic resource conservation – especially as returns from the latter may consist of non-market benefits that accrue to people other than the farmer. This divergence will be further compounded by the existence of distortions in the values of inputs and outputs such that they do not reflect their economic scarcity. The above-described divergence between private and public returns is important. As Pearce and Moran (1994) note, the recognition of the broader total economic value (TEV – see Box 93) of natural assets can be instrumental in altering decisions about their use, particularly in investment decisions that present a clear choice

429

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

Box 93 Economic values Livestock keepers benefit from the conservation of livestock diversity because of their need for animals that are able to produce in diverse agro-ecosystems, and fulfil a range of functions. In addition to supplying products for sale or home consumption, livestock provide input functions related to other farm/ household activities. Livestock provide manure to enhance crop yields, transport for inputs and products, and also serve for traction. Where rural financial and insurance markets are not well developed, they enable farm families to smooth variation in income and consumption levels over time. Livestock constitute savings and insurance, buffering against crop failure and cyclical patterns in crop-related income. They enable families to accumulate capital and diversify, and serve a range of sociocultural roles related to the status and the obligations of their owners (Jahnke, 1982; Anderson, 2003). Livestock also play a role in the maintenance of ecosystems; for example, managed grazing is increasingly viewed as an important tool for conservation. The values mentioned in the above paragraph are components of direct or indirect use value. Other values are not related to use, but simply to the existence of the breeds (existence and bequest values). Another type of value arises from the notion of uncertainty about the future. The latter result from the motivation to avert risk (option value), and from the irreversibility of the loss of a breed and the related loss of information. The “Total Economic Value” (TEV) is formally equal to the sum of all direct and indirect use values plus non-use and option values: TEV = DUV + IUV + OV + BV + XV where:

Direct Use Values (DUV) are the benefits resulting from, inter alia, actual uses, such as for food, fertilizer and hides, as well as cultural/ritual uses. Indirect Use Values (IUV) are the benefits deriving from ecosystem functions. For example, some animals play a key role in the dispersion of certain plant species. Option Values (OV) are derived from the value given to safeguarding an asset for the option of using it at a future date. It is a kind of insurance value (given uncertainty about the future and risk aversion) against the occurrence of, for example, a new animal disease or drought/climate change. Subtly different from, but related to, option values are quasi-option values. The latter relates to the extra value attached to future information made available through the preservation of a resource. Quasi-option values arise from the irreversible nature of breed loss (after which no further learning can take place); they are not related to the risk aversion of the decision makers. Bequest Values (BV) measure the benefit accruing to any individual from the knowledge that others might benefit from a resource in the future; and Existence Values (XV) are derived simply from the satisfaction of knowing that a particular asset exists (e.g. blue whales, capybaras or N´Dama cattle). Some asset values may overlap between these categories, and double counting has to be avoided. Attempts to isolate option, bequest and existence values can be problematic. Underlying principles and procedures for such valuation are still debated.

between erosion/destruction or conservation. When the activity of biodiversity (and genetic resource) conservation generates economic values that are not captured in the market place, the result of this “failure” is a distortion in which the incentives

are against genetic resource conservation, and in favour of the economic activities that erode such resources. Such outcomes are, from an economic viewpoint, associated with market failure (i.e. distortions arising from the “missing markets” in

430

Sources: adapted from Arrow and Fisher (1974); Jahnke, (1982); Pearce and Moran, (1994); Anderson, (2003); Roosen et al., (2005).

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

the external benefits generated by biodiversity conservation); intervention failure (i.e. distortions caused by government actions in intervening in the workings of the market place, even where those appear to serve some social purpose); and/or global appropriation failures (i.e. the absence of markets/mechanisms to capture globally important external values). Note that global missing markets can co-exist with local market failure and intervention failure. The loss of biodiversity and genetic resources is a case in point. It is apparent from the above typology of values that current economic decisions are largely based on the first category, direct use values, although the other categories may be of equal or greater importance. For example, it has been estimated that approximately 80 percent of the value of livestock in low-input developing-country systems can be attributed to non-market roles, while only 20 percent is attributable to direct production outputs. By contrast, over 90 percent of the value of livestock in high-input developed-country production systems is attributable to the latter (Gibson and Pullin, 2005). By focusing exclusively on direct use values, biodiversity and genetic resource conservation are likely to be consistently undervalued, resulting in a bias towards activities that are incompatible with their conservation.

2

Development of methodologies for economic analysis

Although there is a large body of literature on the economic benefits of improved breeds in intensive (largely developed-country) commercial agriculture, the importance of indigenous breeds and trait values in the subsistence production systems typical of developing countries have been much less studied. There is an extensive amount of conceptual and theoretical literature concerning sources of value arising from genetic resources and biodiversity in general (usually referring to plants and wild animals). However, it is only since an FAO/ILRI workshop (ILRI, 1999) identified potential AnGR valuation methodologies, and

subsequent initiatives by ILRI (Economics of AnGR Conservation and Sustainable Use Programme) and its partners to test these methodologies, that significant research into the matter has been carried out. Such tools and their findings have, as yet, rarely been put to use in situations that influence policy-making and farmer livelihoods. Further research is urgently needed to better understand implications for genotype preferences of an increasingly dynamic context characterized by, inter alia: • globalization of markets; • climate change and environmental degradation; • the occurrence of new epidemic animal diseases; • developments in the field of biotechnology; and • policy developments related to the CBD. Global efforts to eradicate poverty, as embodied in the Millennium Development Goals, also require an improved understanding of the potential contributions of alternative genotypes to poverty alleviation, in order to improve pro-poor targeting of AnGR programmes. In this context, research supporting institutional innovations and technology-adoption also play an important role. Such areas are critical for the management of AnGR and have important socioeconomic dimensions. There are a number of reasons for the relatively slow development of the economics of AnGR, including: the fact that the measurement of the benefits of germplasm diversity to livestock development is difficult; the limited availability of the data required to carry out economic analysis; and the importance of considering non-market values of livestock – obtaining such data frequently requires the modification of economic techniques for use in conjunction with participatory and rapid rural appraisal methods. Despite the difficulties, there are a range of analytical techniques from other areas of economics that can be adapted for carrying out such analyses. These methodologies are reviewed

431

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

TABLE 102 Overview of valuation methodologies Valuation methodology

Purpose

Contribution to conservation and sustainable use of AnGR

Group 1: Methodologies for determining the actual economic importance of the breed (mostly of interest to policy makers and breeders, as well as some farmers) Aggregate Demand & Supply

Identify value of breed to society.

Value potential losses associated with AnGR loss.

Cross-sectional Farm and Household

Identify value of breed to society.

Value potential losses associated with AnGR loss.

Aggregated Productivity Model

Determine farmer net returns by breed.

Justify economic importance of given breed in the context of multiple limiting inputs.

IPR and Contracts

Market creation and support for “fair and equitable” sharing of AnGR benefits.

Generate funds and incentives for AnGR conservation.

Contingent Valuation Methodologies I (e.g. dichotomous choice, contingent ranking, choice experiments)

Determine farmer trait value preferences and net returns by breed.

Justify economic importance of given breed.

Market Share I

Indicate current market value of a given breed.

Justify economic importance of given breed.

Group 2: Methodologies for determining the costs and benefits of AnGR conservation programmes and for targeting farmers for participation (mostly of interest to policy makers and farmers) Contingent Valuation Methodologies II (e.g. dichotomous choice, contingent ranking, choice experiments)

Identify society’s willingness to pay (WTP) for the conservation of AnGR. Identify farmer willingness to accept (WTA) compensation for raising indigenous AnGR instead of exotics.

Define maximum economically justified conservation costs.

Production Loss Averted

Indicate magnitude of potential production losses in the absence of AnGR conservation.

Justify conservation programme costs of at least this magnitude.

Opportunity Cost

Identify cost of maintaining AnGR diversity.

Define opportunity cost of AnGR conservation programme.

Market Share II

Indication of current market value of a given breed.

Justify conservation programme costs.

Least Cost

Identify cost-efficient programme for the conservation of AnGR.

Define minimum cost of conservation programme.

Safe Minimum Standard

Assess trade-offs involved in maintaining a minimum viable population.

Define opportunity cost of AnGR conservation programme.

Group 3: Methodologies for priority setting in AnGR breeding programmes (mostly of interest to farmers and breeders) Evaluation of Breeding Programme

Identify net economic benefits of stock improvements.

Maximize economic benefits of conserved AnGR.

Genetic Production Function

Identify net economic benefits of stock improvements.

Maximize expected economic benefits of conserved AnGR.

Hedonic

Identify trait values.

Value potential losses associated with AnGR loss. Understand breed preferences.

Farm Simulation Model

Model improved animal characteristics on farm economics.

Maximize economic benefits of conserved AnGR.

Source: adapted from Drucker et al. (2001).

432

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

by Drucker et al. (2001) who broadly categorize them into three (non-mutually exclusive) groups on the basis of the practical purpose for which they may be used (see Table 102): v group 1) determining the actual economic importance of the breed at risk; v group 2) determining the costs and benefits of AnGR conservation programmes, and targeting farmers for participation; and v group 3) priority setting in AnGR breeding programmes. A number of these methodologies have significant conceptual shortcomings and intensive data requirements (see Drucker et al., 2001 for a detailed description). However, they have been shown to produce useful estimates of the values that are placed on market, non-market and potential breed attributes of the type useful for designing breeding and conservation strategies. The following section presents an overview of the methodologies. The objective is both to show the potential usefulness of the methodologies, as well as to provide information (inevitably location-specific) on the economic importance of indigenous AnGR. To this end, a number of specific studies are presented as illustrative examples of the application of the various tools. Many of the findings give useful insights into the value of particular indigenous livestock breeds within the production systems studied. Salient conclusions are highlighted at the start of each subsection. A more detailed overview can be found in Drucker et al. (2005), and an annotated bibliography of literature in this field is provided by Zambrano et al. (2005).

3

Application of economic methodologies in animal genetic resources management

The following examples are presented in the context of the classification presented in Table 102.

3.1 Value of animal genetic resources to farmers9 • Adaptive traits and non-income functions form important components of the total value of indigenous-breed animals to livestock keepers. • Conventional productivity evaluation criteria are inadequate to evaluate subsistence livestock production and have tended to overestimate the benefits of breed substitution. Tano et al. (2003) and Scarpa et al. (2003a; 2003b) used stated preference choice experiments (CE) to value the phenotypic traits expressed in indigenous breeds of livestock. Adaptive traits and non-income functions are shown to form important components of the total value of the animals to livestock keepers. In the study carried out by Tano et al. (2003) in West Africa, for example, the most important traits for incorporation into the goals of breed improvement programme were found to be disease resistance, fitness for traction, and reproductive performance. Beef and milk production were less important. The results of these studies also show that it is possible to investigate values of genetically determined traits that are currently not widely recognized in livestock populations, but are desirable candidates for breeding or conservation programmes (e.g. disease resistance). Karugia et al. (2001) used an aggregate demand and supply approach covering both national and farm levels. They argue that conventional economic evaluations of crossbreeding programmes have overestimated their benefits by ignoring subsidies, the increased 9

Using Group 1 valuation methodologies (see Table 102).

433

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

costs of management such as veterinary support services, and the higher levels of risk and socioenvironmental costs associated with the loss of the indigenous genotypes. Applied to dairy farming in Kenya, the results suggest that at the nationallevel, cross-breeding has had an overall positive impact on society’s welfare (based on a consumer/ producer surplus measure), although taking important social cost components into account substantially lowers the net benefits. Farm-level performance is, however, little improved under “traditional” production systems by replacing the indigenous Zebu with exotic breeds. Comparing the performance of different genotypes (indigenous goats vs. exotic crosses), Ayalew et al. (2003) come to a similar conclusion. The secondary importance of meat and milk production traits in many production systems leads these authors to argue that conventional criteria for the evaluation of productivity are inadequate for subsistence livestock production systems, because: • they fail to capture non-marketable benefits of the livestock; and • the core concept of a single limiting input is inappropriate to subsistence production, as multiple limiting inputs (livestock, labour, land) are involved in the production process. The study involved the use of an aggregated productivity model to evaluate subsistence goat production in the eastern Ethiopian highlands. The results show that indigenous goat flocks generated significantly higher net benefits under improved than under traditional management, which challenges the prevailing notion that indigenous livestock do not adequately respond to improvements in the level of management. Furthermore, it is shown that under the subsistence mode of production considered, the premise that cross-bred goats are more productive and beneficial than the indigenous goats is wrong. The model, thus, not only underlines the value of indigenous AnGR to farmers, but also provides a more realistic platform upon which to propose sound improvement interventions.

3.2 Costs and benefits of conservation10 • The costs of implementing an in situ breed conservation programme may be relatively small, both when compared to the size of subsidies currently being provided to the commercial livestock sector, and with regard to the benefits of conservation. However, few such conservation initiatives exist, and even where the value of indigenous breeds has been recognized and support mechanisms implemented, significant shortcomings can be identified. • Similar work regarding the costs and benefits of the ex situ (cryo)conservation of livestock remains limited. However, under the assumption that technical feasibility brings the cost of cryoconservation and regeneration of livestock species to within the same level of magnitude as that of plants, extensive conservation efforts would be justified on economic grounds.

In situ conservation Cicia et al. (2003) show that a dichotomous choice stated preference approach can be used to estimate the benefits of establishing a conservation programme for the threatened Italian Pentro horse. A bio-economic model was used to estimate the costs associated with conservation, and a cost–benefit analysis was subsequently realized. Benefit estimates were based on society’s willingness to pay for conservation and, therefore, may be associated, in this particular case, with an existence value. The results not only show a large positive net present value associated with the proposed conservation activity (benefit/ cost ratio > 2.9), but also show that this approach is a useful decision-support tool for policy-makers involved in allocating scarce funds to a growing number of animal breeds facing extinction. A case study of the endangered Box Keken pig breed in Yucatan, Mexico revealed large net present values associated with conservation (Drucker and 10

434

Using Group 2 valuation methodologies (see Table 102).

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

TABLE 103 Conservation benefits and costs under a range of valuation methodologies – the case of the Box Keken pig (Yucatan, Mexico) Valuation methodologies*

Measure of conservation and sustainable use benefits US$ per annum

Market share

US$490 000

Production loss averted (Yucatan State only)

US$1.1 million

Contingent valuation (consumer taste test)

US$1.3 million

Contingent valuation (producer choice experiment) and least cost/ opportunity cost approach

Measure of conservation costs US$ per annum

US$2 500–3 500

Source: Drucker and Anderson (2004). *See Table 102.

Anderson, 2004). Three methodologies for valuing the benefits of conservation and sustainable use of the breed – market share, production loss averted and contingent valuation (consumer taste test) – were tested and critically assessed. The costs of conservation were estimated with the use of contingent valuation (producer choice experiment) and least cost/opportunity cost approaches. A shortcoming of the first two techniques for valuing the benefits is that they are not based on consumer surplus measures, i.e. do not account for price changes and substitution possibilities should breed loss occur. Despite the identified shortcomings, and the fact that values can only be approximated, the study indicates that the benefits of conservation clearly outweigh the costs in this case (Table 103). Even where the value of indigenous breeds has been recognized and support mechanisms implemented, significant failings can be identified. Signorello and Pappalardo (2003), in an examination of livestock biodiversity conservation measures and their potential costs in the EU, report that many breeds at risk of extinction according to the FAO World Watch List are not covered by support payments as they do not appear in countries’ Rural Development Plans. Furthermore, the results show that where payments are made, they do not take into account the different extinction risks faced by the different

breeds. Moreover, payment levels are inadequate, meaning that it can still remain unprofitable to rear indigenous breeds. Ideally, support payments should be set at a level that reflects society’s willingness to pay for conservation, but this is not usually the case and may not always be necessary to ensure profitability. The lack of adequate incentives for the conservation of indigenous breeds is despite the fact that conservation costs have been shown, in a number of case studies described by Drucker (2006), to be relatively small. Drawing on the safe minimum standards (SMS) literature, the framework used in this study assumes that the benefits of indigenous livestock breed conservation can be maintained, as long as a minimum viable population of the breed is maintained. In general, the costs of implementing an SMS are made up of the opportunity cost differential (if any exists) of maintaining the indigenous breed rather than an exotic or cross-breed. In addition, the administrative and technical support costs of the conservation programme also need to be accounted for. Empirical cost estimates were obtained using data from economic case studies (Italy and Mexico), based on an SMS that is equivalent to the FAO measure of “not at risk”, i.e. approximately 1 000 breeding animals. The results support the hypothesis that the costs of implementing an SMS are low (depending on the

435

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

species/breed and location, these ranged from between approximately Euro 3 000 and 425 000 per annum), both when compared with the size of subsidies currently being provided to the livestock sector (less than 1 percent of the total subsidy) and with regard to the benefits of conservation (benefit/cost ratio greater than 2.9). The costs proved to be lowest in the developing country, which is encouraging given that an estimated 70 percent of the livestock breeds existing today are in developing countries, and that this is where the risk of loss is highest (Rege and Gibson, 2003). More extensive quantification of the components required to determine SMS costs nevertheless needs to be undertaken before it can be applied in practice. Such economic valuation needs to cover both the full range of breeds/species being considered, and ensure that as many as possible of the elements making up their total economic value are accounted for.

Ex situ conservation Similar work regarding the costs and benefits of the ex situ (cryo)conservation of livestock remains limited. Cryopreservation technologies for livestock, although advancing rapidly, are still well-developed only for a handful of species. Nevertheless, Gollin and Evenson (2003) argue that assuming that technical feasibility brings the cost of cryoconservation and regeneration of livestock species to within the same level of magnitude as that of plants, “there cannot be much doubt that the economics would justify extensive conservation efforts” (i.e. option values are likely to be much higher than conservation costs).

3.3 Targeting of farmers for participation in in situ breed conservation programmes11 v In situ conservation programmes play a crucial role in the context of AnGR. v Household characteristics play an important role in determining differences in farmers’ breed preferences. This additional information can be of use in designing costeffective conservation programmes. Wollny (2003) argues that community-based management approaches are likely to be required to play an increasingly important role in strategies that aim to improve food security and to alleviate poverty through the conservation of AnGR. This is because the utilization of indigenous livestock populations depends, in large part, on the ability of communities to decide on and implement appropriate breeding strategies. The community-based management of AnGR is also considered to play a critical role in poverty alleviation (FAO, 2003). In the context of crops (Meng 1997), proposed that conservation programmes should target those households that are the most likely to continue to maintain local varieties. As these households will be the least costly to incorporate into a conservation programme, a “least cost” programme can be identified. The cost of an in situ conservation programme can, thus, be expressed as the cost necessary to raise the comparative advantage of such breeds above that of competing breeds, species, or off-farm activities. A relatively small investment may suffice to maintain their advantage in a particular farming system. This conceptual approach to identifying lowcost conservation strategies has recently been applied to estimate conservation costs for creole pigs in Mexico (Scarpa et al., 2003b; Drucker and Anderson, 2004) and Boran cattle in Ethiopia (Zander et al., forthcoming).

11

436

Using Group 2 valuation methodologies (see Table 102).

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

Scarpa et al. (2003b) show that for creole pigs in Mexico, the respondent’s age, years of schooling, size of the household, and the number of economically active members of the household, were important factors in explaining breed trait preferences. Younger, less-educated, and lower-income households placed relatively higher values on the attributes of indigenous piglets compared to exotics and their crosses (Drucker and Anderson, 2004). Pattison’s (2002) findings further corroborate these results. In the context of a ten-year conservation programme that would bring the creole pig population to a sustainable size considered “not at risk” under the FAO classification system; the findings indicate that small, less well-off households would require lower levels of compensation, or even (in 65 percent of cases) no compensation at all. The premise of this set of studies is that continued conservation of genetic resource diversity on-farm makes most economic sense in those locations where both society and the farmers who maintain it benefit the most. Mendelsohn (2003) argues that where there is a divergence between private (farmer) and public values, conservationists must first make the case for why society should be willing to pay to protect apparently “unprofitable” AnGR, and then must design conservation programmes that will effectively protect what society treasures.

3.4 Priority setting in livestock conservation programmes12 • Conservation policy needs to promote cost-efficient strategies, and this can be achieved through the development of “Weitzman-type” decision-support tools. Such tools permit the allocation of a given budget among a set of breeds such that the expected amount of between-breed diversity conserved is maximized.

12

Using Group 2 valuation methodologies (see Table 102).

Simianer et al. (2003) and Reist-Marti et al. (2003) provide one of the few examples of the conceptual development of a decision-support tool in the field of AnGR. Recognizing the large number of indigenous livestock breeds that are currently threatened, and the fact that not all can be saved given limited conservation budgets, a framework is elaborated for the allocation of a given budget among a set of breeds so that the expected amount of between-breed diversity conserved is maximized. Drawing on Weitzman (1993) it is argued that the optimum criterion for a conservation scheme is to maximize the expected total utility of the set of breeds, which is a weighted sum of diversity, extinction probabilities and breed conservation costs (see Section F: 8.2 for further discussion of this approach). Drawing on Group 2 valuation methodologies (see Table 102) is currently postulated as a means of estimating conservation costs. However, Group 1 methodologies could be used should a livelihoods rather than a conservation cost approach be adopted. Both this, and the original Weitzman study, used measures of diversity based on genetic distances. Note, however, that alternative measures of diversity could also be used – for example, measures that include both between and within-breed diversity (Ollivier and Foully, 2005) or those drawing on functional diversity, based on the existence of unique attributes in certain breeds (see Brock and Xepapadeas (2003) for a plant genetic resource illustration). Implications for the choice of breeds for inclusion in conservation programmes may well differ depending on how the diversity index is constructed and the overall goal of the conservation programme (conservation of genetic diversity per se, maximizing the number of unique traits conserved, or maximizing the livelihood contribution of the livestock diversity conserved). Where such models are sufficiently specified and essential data on key parameters are available (currently lacking for conservation costs and benefits or contribution to livelihoods), the framework can be used for rational decisionmaking on a global scale. See section F:8 for further discussion of methods for priority setting in conservation.

437

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

3.5 Priority setting in livestockbreeding strategies13 • Economic analysis has demonstrated the magnitude of the contribution of genetic selection, for example using selection indices, to increased production. • Methods are needed not only to account for the current set of economic objectives, but also to include foreseeable and even unpredictable future needs. • Hedonic approaches14 are useful to evaluate the importance of certain attributes or characteristics to the value of animals or animal products including their influence on selection strategies. Breeding programmes have long used a selection index as a device for multiple-trait selection in livestock. For example, Mitchell et al. (1982) measured the value of genetic contributions to pig improvement in the United Kingdom by determining the heritability of important characteristics, and isolating the genetic contributions to improved performance. Using linear regression techniques to compare control and improved groups over time, they found that the returns were substantial, with costs in the region of £2 million per annum relative to benefits of £100 million per annum. The use of cross-breeding in commercial production was estimated to contribute approximately £16 million per annum. Farm-level simulation models have been built for several species under high-input management, and have also focused on valuing heritable trait gain. Smith (1985), in the context of the importance of accounting for option values in genetic production function models, argues that genetic selection based on the current set of economic objectives is suboptimal in an intertemporal 13

Using Group 3 valuation methodologies (see Table 102).

14

Hedonic approaches are based on the idea that the total value of an animal can be decomposed into the values of individual characteristics. Statistical methods are used to estimate the contribution of each characteristic to the total value based on the market prices paid for animals with different combinations of characteristics.

438

context. Instead, given uncertainty about future needs, selection should be “directed to cater for foreseeable and even unpredictable futures” (Smith, 1985, p. 411). In particular, Smith (1984) advocates the storage of stocks with traits that are currently not economically desirable because of temporary market demands and/or production conditions (e.g. market or grading requirements, carcass or product composition, or special behavioural adaptations to current husbandry conditions). Using hedonic approaches, Jabbar et al. (1998) show that in Nigeria, although there were some differences in prices that were solely because of breed, most variation in prices was because of such variables as wither height and girth circumference that vary from animal to animal within breeds. Variation because of type of animal or month of transaction was also greater than that because of breed. Jabbar and Diedhiou (2003) show that a hedonic approach used to determine livestock keepers’ breeding practices and breed preferences in southwest Nigeria, confirms a strong trend away from trypanotolerant breeds. Richards and Jeffrey (1995) identified the value of relevant production and type traits for dairy bulls in Alberta, Canada. A hedonic valuation model was estimated, which modelled semen price as a function of individual production and longevity characteristics for a sample of Holstein-Friesian bulls.

3.6 General policy analysis15 The current rapid rate of loss of AnGR diversity is the result of a number of underlying factors. While, in some cases, changes in production systems and consumer preferences reflect the natural evolution of developing economies and markets, in other cases, production systems, breed 15

Potentially using Group 2, as well as Group 1 valuation methodologies (see Table 102).

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

choice and consumer preferences have been distorted by local, national and international policy. Such distortions may arise from macroeconomic interventions (e.g. exchange and interest rates); regulatory and pricing policy (e.g. taxation, price controls, market and trade regulations); investment policy (e.g. infrastructure development); and institutional policy (e.g. land ownership and genetic resource property rights). While the impact of policy factors on AnGR is readily discernable in broad terms, little is known about their relative importance.

4

Implications for policies and future research

The above studies reveal not only that there are a range of methodologies that can be used to value livestock keeper breed/trait preferences, but that they can be of use in designing policies that counter the present trend towards marginalization of indigenous breeds. In particular, it becomes possible to, inter alia (Drucker and Anderson, 2004): • recognize the importance that livestock keepers place on adaptive traits and nonincome functions, and the need to consider these in breeding programme design; • identify those breeds that are a priority for participation in cost-efficient diversitymaximizing conservation programmes; and • contrast the costs involved with the large benefits non-livestock keepers place on breed conservation. Nevertheless, as recent advances in economic valuation for livestock genetic resources have eased some (but by no means all) methodological/ analytical constraints, the issue of data availability has become relatively more critical. Data requirements imply the need to inter alia: • measure breed performance parameters; • characterize actual and potential breeding systems; • identify uses and farmers’ trait preferences

(including eliciting the values that farmers place on specific market/non-market traits and the trade-offs they are willing to make between traits) for local breeds under different production systems, as well as the forces influencing such factors and the use of alternative breeds; • identify factors affecting livestock demand and prices, including the impact of policyinduced changes in agricultural commodity (e.g. forage/crop) prices and external (e.g. veterinary) input costs in the context of different breed use; • carry out ex ante analysis of the effects on livelihoods of using alternative breeds, together with constraints to adoption and potential access/dissemination mechanisms; • consider the role of such factors as land tenure, agricultural potential, population density, market access and integration, licensing requirements, tax regimes, credit and extension programmes and education; and • improve understanding of the importance of continued access and trade in livestock germplasm for research and development purposes, together with the nature of the costs and benefits arising from AnGR research. Despite a wealth of livestock production data at the national level, such information tends to be limited to the principal breeds and largely ignores important non-market contributions. Information on local breeds in developing countries is extremely limited. Initiatives such as FAO’s DAD-IS and ILRI’s DAGRIS systems are supporting nationallevel programmes. The challenge is now to raise awareness regarding the important role of economic analysis in improving farm AnGR conservation and sustainable use. National capacities must also be strengthened in order to enable the application of the relevant methodologies/decision-support tools, and to integrate them into the wider national livestock development process. In this

439

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

way, further work on the economics of AnGR (including in dynamic systems evolution contexts and integrated with other components of agrobiodiversity), and the subsequent design of appropriate incentive mechanisms, can be applied in contexts where the results can be taken up so as to actively benefit farmers and support the work of national researchers and policy-makers.

References16 Anderson, S. 2003. Animal genetic resources and sustainable livelihoods. Ecological Economics, 45(3): 331–339. Arrow, K.J. & Fisher, A.C. 1974. Environmental preservation, uncertainty, and irreversibility. Quarterly Journal of Economics, 88(2): 312–319. Ayalew, W., King, J.M., Bruns, E. & Rischkowsky, B. 2003. Economic evaluation of smallholder subsistence livestock production: lessons from an Ethiopian goat development program. Ecological Economics, 45(3): 473–485. Brock, W. & Xepapadeas, A. 2003. Valuing biodiversity from an economic perspective: a unified economic, ecological and genetic approach. American Economic Review, 93(5): 1597–1614. Cicia, G., D’Ercole, E. & Marino, D. 2003. Costs and benefits of preserving farm animal genetic resources from extinction: CVM and bio-economic model for valuing a conservation program for the Italian Pentro horse. Ecological Economics, 45(3): 445–459. Drucker, A.G. 2006. An application of the use of safe minimum standards in the conservation of livestock biodiversity. Environment and Development Economics, 11(1): 77–94. Drucker A.G. & Anderson, S. 2004. Economic analysis of animal genetic resources and the use of rural appraisal methods: Lessons from South-East Mexico. International Journal of Sustainable Agriculture, 2(2): 77–97. Drucker, A.G., Gómez, V. & Anderson, S. 2001. The economic valuation of farm animal genetic resources: a survey of available methods. Ecological Economics, 36(1): 1–18. Drucker, A.G., Smale, M. & Zambrano, P. 2005. Valuation and sustainable management of crop and livestock biodiversity: a review of applied economics literature. SGRP/IFPRI/ILRI. (available at www.ilri.org). 16 See www.ilri.org for full text versions of a number of these papers.

440

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

FAO. 2003. Community-based management of animal genetic resources. Proceedings of the workshop held in Mbabane, Swaziland, 7–11 May 2001. FAO/ SADC/UNDP/GTZ/CTA. Rome.

Ollivier, L. & Foulley, J. 2005. Aggregate diversity: new approach combining within- and between-breed diversity. Livestock Production Science, 95(3): 247-254.

Gibson, J.P. & Pullin, R.S.V. 2005. Conservation of livestock and fish genetic resources. Rome. CGIAR Science Council Secretariat.

Pattison, J. 2002. Characterising backyard pig keeping households of rural Mexico and their willingness to accept compensation for maintaining the indigenous Creole breed: A Study of Incentive Measures and Conservation Options. University of London. (MSc thesis).

Gollin, D & Evenson, R. 2003. Valuing animal genetic resources: lessons from plant genetic resources. Ecological Economics, 45(3): 353–363. ILRI. 1999. Economic valuation of animal genetic resources. Proceedings of an FAO/ILRI workshop held at FAO Headquarters, Rome, Italy, 15–17 March 1999. Nairobi. International Livestock Research Institute. Jabbar, M.A. & Diedhiou, M.L. 2003. Does breed matter to cattle farmers and buyers? Evidence from West Africa. Ecological Economics, 45(3): 461–472. Jabbar, M.A., Swallow, B.M., d’Ieteren, G.D.M. & Busari, A.A. 1998. Farmer preferences and market values of cattle breeds of west and central Africa. Journal of Sustainable Agriculture, 12: 21–47. Jahnke, H.E. 1982. Livestock production systems and livestock development in Tropical Africa. Kiel, Germany. Kieler Wissenschaftsverlag Vauk. Karugia, J., Mwai, O., Kaitho, R., Drucker, A., Wollny, C. & Rege, J.E.O. 2001. Economic analysis of crossbreeding programmes in sub-Saharan Africa: a conceptual framework and Kenyan case study. Animal Genetic Resources Research 2. Nairobi. International Livestock Research Institute. Mendelsohn, R. 2003. The challenge of conserving indigenous domesticated animals. Ecological Economics, 45(3): 501–510. Meng, E.C.H. 1997. Land allocation decisions and in situ conservation of crop genetic resources: The case of wheat landraces in Turkey. University of California, Davis, California, USA. (PhD thesis) Mitchell, G., Smith, C., Makower, M. & Bird, P.J.W.N. 1982. An economic appraisal of pig improvement in Great Britain. 1. Genetic and production aspects. Animal Production, 35(2): 215–224.

Pearce, D. & Moran, D. 1994. The economic value of biodiversity. London. Earthscan. Rege, J.E.O. & Gibson, J.P. 2003. Animal genetic resources and economic development: issues in relation to economic valuation. Ecological Economics, 45(3): 319-330. Reist-Marti, S., Simianer, H., Gibson, G., Hanotte, O. & Rege, J.E.O. 2003. Weitzman’s approach and breed diversity conservation: an application to African cattle breeds. Conservation Biology, 17(5): 1299–1311. Richards, T. & Jeffrey, S. 1995. Hedonic pricing of dairy bulls – an alternative index of genetic merit. Department of Rural Economy. Project Report 95–04. Faculty of Agriculture, Forestry, and Home Economics. Edmonton, Canada. University of Alberta Edmonton. Roosen, J., Fadlaoui, A. & Bertaglia. M. 2005. Economic evaluation for conservation of farm animal genetic resources. Journal of Animal Breeding and Genetics, 122(4): 217–228. Scarpa, R., Drucker, A.G., Anderson, S., FerraesEhuan, N., Gómez, V., Risopatrón, C.R. & RubioLeonel, O. 2003a. Valuing genetic resources in peasant economies: the case of ‘hairless’ Creole pigs in Yucatan. Ecological Economics, 45(3): 427-443. Scarpa, R., Ruto, E.S.K., Kristjanson, P., Radeny, M., Drucker, A.G. & Rege, J.E.O. 2003b. Valuing indigenous cattle breeds in Kenya: an empirical comparison of stated and revealed preference value estimates. Ecological Economics, 45(3): 409–426.

441

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

Signorello, G. & Pappalardo, G. 2003. Domestic animal biodiversity conservation: a case study of rural development plans in the European Union. Ecological Economics, 45(3): 487–499. Simianer, H., Marti, S.B., Gibson, J., Hanotte, O. & Rege, J.E.O. 2003. An approach to the optimal allocation of conservation funds to minimise loss of genetic diversity between livestock breeds. Ecological Economics, 45(3): 377–392. Smith, C. 1984. Genetic aspects of conservation in farm livestock. Livestock Production Science, 11(1): 37–48. Smith, C. 1985. Scope for selecting many breeding stocks of possible economic value in the future. Animal Production, 41: 403–412. Swanson, T. 1997. Global action for biodiversity. London. Earthscan. Tano, K., Kamuanga, M., Faminow, M.D. & Swallow, B. 2003. Using conjoint analysis to estimate farmer’s preferences for cattle traits in West Africa. Ecological Economics, 45(3): 393–407. Weitzman, M.L. 1993. What to preserve? An application of diversity theory to crane conservation. The Quarterly Journal of Economics, 108(1): 157–183. Wollny, C. 2003. The need to conserve farm animal genetic resources through community based management in Africa: should policy-makers be concerned? Ecological Economics, 45(3): 341–351. Zander, K., Drucker, A.G., Holm-Muller, K. & Mburu, J. (forthcoming). Costs and constraints of conserving animal genetic resources: the case of Borana cattle in Ethiopia. Zambrano, P., Smale, M. & Drucker, A.G. 2005. A selected bibliography of economics literature about valuing crop and livestock components of agricultural biodiversity. SGRP/IFPRI/ILRI.

442

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

Section F

Methods for conservation 1

Introduction

Breed development is a dynamic process of genetic change driven by environmental conditions and selection by humans, the latter being shaped by the culture and the economic situation. The fact that ecosystems are dynamic and complex and that human preferences change, has resulted in the evolution of breeds and, until recently, a net increase in diversity over time. However, in the past 100 years there has been a net loss of diversity resulting from an increase in the rate of extinction of breeds and varieties. In Europe and the Caucasus alone, 481 mammalian and 39 avian breeds have already become extinct, and another 624 mammalian and 481 avian breeds are at risk. Losses have been accelerated by rapid intensification of livestock production, a failure to evaluate local breeds, and inappropriate breed replacement or cross-breeding facilitated by the availability of high-performing breeds and reproductive biotechnologies (Box 95).

Box 94 Glossary: conservation For the purpose of this report, the following definitions are used: Conservation of animal genetic resources: refers to all human activities including strategies, plans, policies and actions undertaken to ensure that the diversity of animal genetic resources being maintained to contribute to food and agricultural production and productivity, or to maintain other values of these resources (ecological, cultural) now and in the future.

In situ conservation: refers to conservation of livestock through continued use by livestock keepers in the production system in which the livestock evolved or are now normally found and bred. Ex situ in vivo conservation: refers to conservation through maintenance of live animal populations not kept under normal management conditions (e.g. zoological parks and in some cases governmental farms) and/or outside of the area in which they evolved or are now normally found. There is often no clear boundary between in situ and ex situ in vivo conservation and care must be taken to describe the conservation objectives and the nature of the conservation in each case.

Ex situ in vitro conservation: refers to conservation external to the living animal in an artificial environment, under cryogenic conditions including, inter alia, the cryoconservation of embryos, semen, oocytes, somatic cells or tissues having the potential to reconstitute live animals (including animals for gene introgression and synthetic breeds) at a later date.

443

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

Box 95 Red Maasai sheep – accelerating threats The Red Maasai, renowned for its hardiness and disease resistance, especially its resistance to gastrointestinal parasites, is predominantly kept by Maasai pastoralists, as well as by the neighbouring tribes in the semi-arid regions of Kenya and the United Republic of Tanzania. A number of research projects have demonstrated the breed’s resistance to diseases, and high productivity under extremely challenging environments, where other breeds, such as the introduced Dorper perform very poorly. Until the mid-1970s, pure-bred Red Maasai were ubiquitous throughout the pastoral lands of Kenya, probably numbering several million head. In the mid1970s, a subsidized dissemination programme for Dorper rams was established in Kenya. Widespread indiscriminate cross-breeding followed. No instruction was provided to farmers about how to maintain a continuous cross-breeding programme and many farmers continued crossing their flocks to Dorpers, which subsequently proved unsuitable in many production areas. In 1992, and again more recently, the International Livestock Research Institute undertook an extensive search in Kenya and northern parts of the United Republic of Tanzania, but was only able to locate a very small number of purebred animals. The Institute was able to establish a small “pure-bred” flock, but this flock later showed some levels of genetic contamination. The Red Maasai breed is clearly threatened, but the livestock databases DAD-IS and DAGRIS do not identify the breed as threatened, and the breed does not appear in the World Watch List (FAO/UNEP 2000). This is related to the current inability of the systems to document the dilution of breeds. Provided by John Gibson.

While the loss of livestock genetic diversity has greatly increased in recent decades, the extent of the problem has still not been fully evaluated. Information on AnGR provided by FAO member

444

countries is made available to the public in the DAD-IS database. Although a specific call for information on extinct breeds was made in 1999 before compiling the third edition of the World Watch List (FAO/UNEP, 2000), the lists of extinct breeds are probably not complete – uncharacterized local populations in rapidly developing regions of the world may have disappeared without being recorded. Reasons for extinction are either not documented or not readily accessible, and therefore have not been thoroughly analysed. The risk status of many breeds can only be estimated, as breed population census data are often missing or unreliable. The lack of knowledge hinders concerted actions and the setting of conservation priorities.

2

Arguments for conservation

The ratification of the CBD by 188 states indicates a growing international commitment to sustain and protect biodiversity. The CBD calls for conservation and sustainable use of all components of biological diversity including those used for agriculture and forestry. Recognizing the importance of genetic level diversity it provides a mandate to conserve genetic resources for food and agriculture. Article 2 specifically recognizes “domesticated and cultivated species” as an important component of global biological diversity. However, it has been noted that “while a significant international consensus regarding policy has apparently emerged, this consensus is not grounded in a consensually accepted value theory to explain why biodiversity protection, however strongly supported, should be a top priority of environmental policy” (Norton, 2000 in FAO, 2003, p. 105). For example, the argument for maintaining biological diversity for its own sake can be contrasted with the view that in the absence of a clear case for the utility of a breed, its loss should not be of much concern. This chapter presents an

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

overview of the different lines of argument put forward in favour of conservation. The rationale of a conservation programme may include a combination of the following arguments:

2.1 Arguments related to the past Livestock breeds reflect the cultural and historical identity of the communities that developed them, and have been an integral part of the livelihood and traditions of many societies. Loss of typical breeds, therefore, means a loss of cultural identity for the communities concerned, and the loss of part of the heritage of humanity. A further argument relates to the fact that breed development, especially in species with longer generation intervals, will often have involved considerable investments in terms of time, financial expenditure and/or institutional resources. Moreover, historical processes may have given rise to unique outputs that could not easily be recreated. According to this point of view, the decision to abandon such breeds should, therefore, not be taken lightly. There is also a historical dimension to the development of adaptive traits – the longer an animal population has been exposed to an environmental challenge, the greater the possibility that specific adaptive traits have evolved. Areas with climatic extremes or particular disease conditions have given rise to genetically adapted and unique local stocks. These breeds have co-evolved with a particular environment and farming system, and represent an accumulation of both genetic stock, and associated husbandry practices and local knowledge.

2.2 Safeguarding for future needs “Predicting the future is a risky business at best, particularly where human activities are involved” (Clark, 1995 in Tisdell, 2003, p. 369). It is notoriously difficult to predict the future, and people’s expectations are highly diverse. Very negative expectations may at times be more related to unsubstantiated fears than to rational arguments. However, a strong case for concern

about the loss of AnGR diversity can be put forward: “From a long-term point of view, it is possible that concentration on high yielding environmentally sensitive breeds will create a serious problem for the sustainability of livestock production ... it is possible that farmers will lose their ability to manipulate natural environmental conditions. If all environmentally tolerant breeds are lost in the interim, the level of livestock production could collapse.” (Tisdell, 2003, p. 373). Unforeseen developments may be brought about by changes in the ecosystem, in market demands and associated regulations, by changes in the availability of external inputs, by emerging disease challenges, or by a combination of these factors. Global climate change and the evolution of resistance in pathogens and parasites to chemical control are almost certain to affect future livestock production systems, though the nature of the changes remains unclear (FAO, 1992). The possibility of catastrophic losses of AnGR resulting from major disease epidemics, war, bioterrorism or civil unrest, indicates a need to have a secure reserve, such as a genebank, for breeds that are of great economic importance at present. The uncertainty of future needs, in combination with the irreversible nature of events such as species or breed extinction, highlights the need to safeguard the option value17 of diversity. Examples of previously unforeseen needs include the trend among developed-world animal breeders away from production-oriented genetic improvement to focus more on adaptation, disease resistance and feed efficiency. In some developed countries, the importance of conservation grazing has reached an extent that few would have foreseen forty years ago when rare breeds began to be used for this purpose. In the United

17

The option value of diversity is the value given to safeguarding an asset for the option of using it at a future date.

445

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

Box 96 Lleyn sheep of Wales – revival in fortunes in tune with modern demands In the course of the last half century the Lleyn sheep breed of northwest Wales has progressed from the brink of extinction to a breed of widespread national importance in the British sheep industry. Following the Second World War, the breed retreated from the considerable local importance that it had in the first half of the century, and by the 1960s there were a mere seven pure-bred flocks and 500 ewes. In contrast, by 2006 the number of pure breeders exceeds 1 000 spread throughout the United Kingdom, and regional Society sales involve the annual trading of many thousands of Lleyn sheep. This revival was achieved through the determination and enthusiasm of an initially small group of twelve local breeders and supportive advisers. They set up a breed society in 1970 to coordinate breeding policy, register pure-bred flocks and grade up cross-bred sheep (by repeated backcrossing using Lleyn rams). The chief attributes of the breed from the start were its medium size, mothering ability (in its hey-day it was milked after weaning the lamb) and prolificacy, as well as meat and wool quality. An added attraction for flock biosecurity was the suitability of the Lleyn for “closed flock” operations in which the only animals purchased are top-quality rams. These attributes were intensified by organized breeding, partly through the operation of a New Zealand-type nucleus group breeding scheme, involving objective recording (Meat and Livestock Commission) and fast generation turnover. The resulting wide appeal of easily handled ewes, convenient for large and small flock owners, coupled with efficient utilization of expensive land, was

Kingdom, over 600 conservation sites are grazed (although not all with rare or traditional breeds) and as many as 1 000 sites would benefit from such grazing (Small, 2004). Specific breeds which were once under threat but have now proved to be of economic importance include the Piétrain pig. This very lean breed, which is now used in

446

fostered by the support of the Breed Society. This involved shrewd marketing, with well-organized breed sales and information provision for prospective buyers and member breeders. Another important element, as the breed rapidly extended its geographical coverage, was the encouragement given to local devolution. Groups or clubs have been formed on a countrywide basis, currently seven clubs in all, although the parent breed society has maintained its coordinating role and its link with the home base in northwest Wales.

Provided by J B Owen. For further information on the breed see: http://www.lleynsheep.com

Photo credit: David Cragg

a large number of commercial cross-breeding programmes, was hardly known outside the Brabant province of Belgium prior to 1950. It almost became extinct during the Second World War when fat animals were in demand (Vergotte de Lantsheere et al., 1974). Another example is the Lleyn sheep breed from Wales, which

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

during the 1960s was in serious decline and had a population size of only 500 pure-bred ewes (Box 96). The breed has become increasingly popular among sheep farmers in the United Kingdom in recent years and its population has grown to over 230 000. The Wiltshire Horn, another British breed that was once in decline, is also attracting interest because of changing market conditions. The breed sheds its wool – a desirable characteristic when shearing costs can exceed the price obtained for the fleece. Opportunities provided by future developments in biotechnology also need to be considered. Emerging reproductive and genetic technologies already provide greatly increased opportunities to identify and utilize the genetic variation of AnGR, and such technologies are expected to show major advances in future. If diverse AnGR remain available, such technologies should make it possible for developing countries to close the productivity gap with developed countries by selectively combining the best features of different breeds. It is widely accepted that the future option value of AnGR provides a strong reason for conserving AnGR. It is reasonable to assume that changing circumstances and rapidly advancing technologies will require the use of conserved AnGR in the future.

2.3 Arguments related to the present situation The importance of maintaining threatened AnGR does not necessarily relate only to their potential future use under changed circumstances. There are a number of reasons why the use of these resources may be sub-optimal at present. These reasons fall into three main categories: deficits in information, market failures and policy distortions (Mendelsohn, 2003). There are large gaps in knowledge regarding the characteristics of local breeds and their traits or genes that may be important for production, research purposes or to meet other human needs (Oldenbroek, 1999). Imperfect information may lead to the overestimation of the performance of a breed

within a particular production environment where its introduction is being considered, and hence an inappropriate decision regarding its adoption. It is, of course, also possible that imperfect information could lead to farmers unnecessarily retaining their indigenous breed and not adopting alternative breeds that would improve their livelihoods. Policy distortions can put less intensive production systems at a disadvantage and provide disincentives for efficient resource allocation. A narrow focus on high-output breeds may be favoured by policies such as subsidized grain imports, free or subsidized support services (e.g. AI) or support prices for livestock products, which stimulate the intensification processes. For example, in some rapidly industrializing Asian countries important capital subsidies have clearly favoured an industrial mode of development; cheap capital has led to investments in large commercial units associated with high input use and uniform products. Furthermore, development or emergency programmes sometimes promote exotic breeds from donor countries. Finally, political instability and policies unfavourable to vulnerable livestock keeping populations may inhibit the efficient use of AnGR (Tisdell, 2003). Markets may not accurately represent external costs or benefits. Examples of external costs include negative environmental impacts, and undesirable effects on income distribution and equity. External benefits associated with certain breeds may, for example, include their contribution to landscape management. Mendelsohn (2003, p. 10) suggests that: “Conservationists must focus on what the market will not do. They must identify and quantify the potential social benefits of AnGR that have been abandoned by the market.” The preservation of diversity, including withinbreed diversity, serves to maintain stability in production systems. Diverse populations show greater ability to survive, produce and reproduce under conditions of fluctuating feed resources and water supply; extremes of temperature, humidity and other climatic factors; and low levels

447

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

of management (FAO, 1992). There is evidence that they are also less susceptible to catastrophic epidemics (Springbett et al., 2003). In general, genetically uniform populations are less able to respond to strong selection pressures resulting from environmental changes. Maintaining breed diversity enables people to exploit diverse ecological or economic niches. This is particularly the case in marginal and environmentally fragile areas, such as drylands, where most livestock kept by poor farmers are located, and which are characterized by great diversity and high levels of risk. Arguments for existence and bequest values for AnGR,18 remove the need to identify tangible or non-tangible benefits as a justification for conservation. “Biological diversity has intrinsic value and should be conserved for its own sake to the maximum extent possible, regardless of whether any given component can be shown to produce tangible economic benefits” (FAO, 2003, p. 104). However, the development of breeds within domesticated species is primarily the product of human intervention to meet human objectives and values. The argument that the current diversity should be preserved on the grounds of its existence value is, therefore, perhaps more difficult to defend than in the case of the biodiversity of natural ecosystems. Arguments and capacities for conservation vary from region to region. In Western societies, traditions and cultural values are important driving forces, which ensure the development of conservation measures for rare breeds and promote the emergence of niche markets for livestock products. By contrast, in the developing world, the immediate concerns are for food security and economic development. However, most developing countries are already in a process 18

The existence value is derived from the satisfaction of knowing that a particular asset exists; a bequest value is the benefit accruing to any individual from the knowledge that others might benefit from the resource in the future.

448

of economic evolution, and their economies can be expected to become sufficiently developed to support conservation based on cultural heritage and other such drivers at some point in the future. There is a need to ensure that AnGR are not lost before this self-supporting stage is reached.

3

The unit of conservation

A critical first step in the design of AnGR conservation programmes is to decide what is to be conserved. At the molecular genetic level, the genetic diversity present within a livestock species is a reflection of allelic diversity (i.e. differences in DNA sequences) across the 25 000 or so genes (i.e. functional DNA regions) affecting animal development and performance. Conceptually, therefore, the most basic unit of conservation is the allele. An objective might be to design conservation programmes that will both allow maintenance of a preponderance of the alleles that are currently present within a species, as well as providing for the normal accumulation and potential retention of the newly arising mutant alleles which are the fuel for continued animal evolution and improvement. Allelic diversity could, in theory, be quantified by enumeration of the number and frequencies of the various alleles, but for the moment this is an impossible task. In defining the unit of conservation, it must further be recognized that alleles do not act in isolation, and that animal performance in most cases is properly viewed as a result of the interactions of alleles present across the genome. Thus, the process of genetic resource development involves the creation of allelic combinations that support specific desired levels of animal performance and adaptation. Efficient genetic resource conservation, therefore, involves the creation of structures that allow for maintenance of existing genetic combinations of known adaptive or productive value, and for easy access to these combinations to support future animal production needs.

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

Existing livestock breeds are less genetically uniform than most varieties of crop plants, but nonetheless represent the realization of a diverse set of adaptive processes. The population structure of the major livestock species up to the mid-twentieth century conformed closely to the population structure predicted to maximize evolutionary potential. There were many partially isolated subpopulations (the breeds), maintained under diverse conditions, but with periodic exchange of animals among populations and periodic recombination of breeds to yield new genetic combinations. Thus, adoption of the breed as the unit of conservation is expected to maximize the maintenance of evolutionary potential within livestock species, and likewise to maximize access to a broad array of allelic combinations.

4

Conservation of plant versus animal genetic resources

Organization and implementation of the SoWAnGR assessment process was based on the lessons learnt from the global assessment of plant genetic resources (PGR) and the resulting Report on the State of the World’s Plant Genetic Resources (FAO, 1998a). Accordingly, the SoWAnGR process focused on both the preparation of the first Report, and the initiation of actions at national level arising from the process of Country Report preparation. Nevertheless, approaches for conservation of PGR cannot be directly applied to AnGR. In traditional production systems, plant and animal genetic resources are used in comparable ways. Locally adapted breeds and varieties predominate; seed for planting, and breeding animals are drawn from the farmers’ fields, herds and flocks, and genetic diversity within resulting landraces is substantial. Most breeding and development activities are “participatory” (FAO, 1998a) in the sense that decisions regarding the seeds to save for planting and the animals to retain for breeding are made by farmers rather

than professional plant and animal breeders. However, intensification of agriculture has resulted in important changes in patterns of genetic resource utilization and development. In plants, intensification of crop production has generally been accompanied by emergence of a strongly institutionalized and centralized seed production sector dominated by publicly funded national and international centres, and private firms. In contrast, the intensification of the livestock sector is currently much less advanced, and has been a result of, rather than a prerequisite for, economic development. The animal breeding sector is far less centralized and institutionalized than the plant seed sector, although there has been substantial movement towards centralization in the poultry, pig and, to a more limited extent, dairy cattle sectors. Direct involvement of farmers in animal breeding remains substantial for the other livestock sectors, and AnGR utilization and further development remains strongly “participatory” in certain production environments. The different structures of the seed and seedstock sectors in plants and animals have important implications for the conservation of global genetic resources. Table 104 compares a number of biological, operational, and institutional factors that influence conservation activities in plants and animals. Biological differences clearly require different approaches to conservation, but perhaps the most significant difference between the crop and livestock sectors involves institutional capacity for genetic resource management. Many of the institutions of the seed sector already maintain extensive collections of PGR, and actively contribute to the development and release of plant varieties. The databases of the World Information and Early Warning System on Plant Genetic Resources (WIEWS) record the location of over 5.5 million PGR accessions, in some 1 410 ex situ collections around the world (FAO, 2004). Establishing a genebank for animals involves long-term storage of gametes, embryos or somatic cells in liquid nitrogen. Technical aspects of such

449

THE STATE OF THE WO RLD' S AN IMAL GENE T I C RESOURCES FOR FOOD AND AG RICULTURE

PART 4

in vitro conservation in animals are discussed in detail below, but costs to collect, cryoconserve and subsequently reconstitute animal germplasm are many times greater per preserved genome than costs to collect, store and subsequently utilize seeds. Moreover, funding to support the conservation of animal germplasm has been insufficient. As a result, AnGR conservation has much more heavily emphasized in situ approaches. However, with the exception of a small number of developed countries, there has been little action to establish in situ conservation programmes, and the long-term sustainability of such schemes remains uncertain. DAD-IS lists 4 956 extant mammalian breeds and 1 970 extant avian breeds. Few of these are well represented in in vitro collections and almost none have been sampled at levels consistent with FAO (1998b) guidelines for in vitro sampling. Very substantial resources would be required to develop in vitro collections of even the most endangered of these nearly 7 000 livestock breeds. For example, the FAO (1998b) Guidelines for Management of Small Populations at Risk recommend collection of frozen semen from at least 25 males per breed, and use of semen from these males on an additional 25 females per breed to produce frozen embryos. For cattle, with 300 endangered breeds, cryoconservation of semen from 7 500 males and approximately 100 000 embryos would be required. Policy guidelines for ownership, use and management of in vitro collections are yet to be developed. Institutional capacity for AnGR conservation is limited, with only a few national ex situ collections existing, mainly in developed countries. Among the institutions of the Consultative Group on International Agricultural Research (CGIAR), only the International Livestock Research Institute (ILRI) and the International Center for Agricultural Research in the Dry Areas (ICARDA) actively address issues of better management of AnGR, and neither institution has an active programme for long-term storage of germplasm. Ownership of AnGR resides almost exclusively in the private

450

sector. A substantial enhancement of global capacity for conservation and better use of AnGR, with new institutional models and collaboration among public institutions and between public institutions and private farmers, may therefore be required if the recommendations of the SoWAnGR process are to be implemented.

5

Information for conservation decisions

Setting priorities for AnGR conservation requires a process that enables the identification of breeds that contribute most to global genetic diversity and have the greatest potential to contribute to efficient future utilization and further development of that diversity. Additional criteria, such as cultural or heritage values of a breed, will also affect priorities for conservation. Assessing the likely genetic diversity present in a set of breeds may be based on a variety of criteria, including: • trait diversity, which is diversity in the recognizable combinations of phenotypic characteristics that define breed identity; • molecular genetic diversity, based on objective measurements of genetic relationships among breeds at the DNA level; and • evidence for past genetic isolation as a result of either geographical isolation or of breeding policies and cultural preferences applied in the communities where the breeds were developed. Trait diversity is based on heritable phenotypic differences among breeds. When breeds are compared under comparable environmental conditions, trait diversity is necessarily indicative of underlying functional genetic diversity. For this reason, breeds that possess unique or distinctive trait combinations should be given high priority for conservation, because their unique phenotypic characteristics necessarily reflect unique underlying genetic combinations.

STATE OF THE A R T I N THE MA N AGEMEN T OF A N I MAL GENE T I C RESOU RC ES

TABLE 104 Comparisons of biological, operational and institutional factors influencing plant and animal genetic resources conservation Factor

Plants

Animals

Economic value of production per individual

Low to very low

Moderate to high

Reproductive rate (number of progeny per individual per generation)

High to very high (1000s)

Very low (